Density-Functional Theory PDF

You might also like

Download as pdf
Download as pdf
You are on page 1of 69
cond-mat/0211443v5 [cond-mat.mtrl-sci] 18 Nov 2006 arXiv A Bird’s-Eye View of Density-Functional Theory Klaus Capelle Departamento de Fésiea ¢ Informatica Instituto de Fisica de Séo Carlos Universidade de Sto Paulo Caixa Postal 369, Séo Carlos, 13560-970 SP, Brazil keywords: density-functional theory, electronic-structure theory, electron cor- relation, many-body theory, local-density approximation Abstract ‘This paper is the outgrowth of lectures the author gave at the Physies Institute and the Chemistry Institute of the University of Sio Paulo at Sio Carlos, Brazil, and at the VIII'th Summer School on Electronic Structure of the Brazilian Physical Society. It is an attempt to introduce density-funetional theory (DFT) in a language accessi- ble for students entering the field or researchers from other fields. Tt is not meant to be a scholarly review of DFT, but rather an infor- ‘mal guide to its conceptual basis and some recent developments and ‘advances, The Hohenberg-Kohn theorem and the Kohn-Shamn equa- tions aro discussed in some detail. Approximate density functionals, selected aspects of applications of DFT, and a variety of extensions of standard DFT are also discussed, albeit in less detail, Through- out it is attempted to provide a balanced treatment of aspects that are relevant for chemistry and aspects relevant for physics, but. with ‘a strong bins towards conceptual foundations. The paper is intended to be read before (or in parallel with) one of the many excellent more echnical veviews available in the literature. Contents 6 Preface What is density-functional theory? DFT as a many-body theory 3.1 Punetionals and their derivatives 3.2 ‘The Hohenberg-Kohn theorem 4 8 8 10 3.3 Complications: NV and v-representability of densities, and nonunique- ness of potentials . . 34 A proview of practical DPT. 3.5 From wave functions to density functionals via Green’ s fune- tions and density matrices 3.5.1 Green’s functions - 3.5.2 Density matrices. DFT as an effective single-body theory: The Kohn-Sham ‘equations 4.1 Exchange-correlation energy: definition, interpretation. and exact properties 4.1.1 Exchange-correlation energy. 4.1.2 Different perspectives on the correlation energy 4.1.3 Exact properties 4.2 Kobn-Sharn equations 42.1 Detivation of the Kobn-Sham equations 4.2.2. "Phe eigenvalues of the Kohn-Sham equation 4.2.3 Hartree, Hartree-Fock and Dyson equations 4.3 Basis functions : ‘Making DFT practical: Approximations 5.1 Local functionals: LDA . 5.2 Semilocal functionals: GEA, GGA and beyond . 5.3 Orbital functionals and other nonlocal approximations: hy- brids, Meta-GGA, SIC, OEP, etc, : Extensions of DFT: New frontiers and old problems References 15 16 19 19 21 25 26 26 28 30 32 32. 35, 37. 39 aL 43, 45 49 53 59 1 Preface ‘This paper is the outgrowth of lectures the author gave at the Physics Institute and the Chemistry Institute of the University of Séo Paulo at Siio Carlos, Brazil, and at the VIIPth Summer School on Electronic Structure of the Brazilian Physical Society [1]. The main text is a description of density- functional theory (DFT) at a level that should be accessible for students entering the field or researchers from other fields. A large number of footnotes provides additional comments and explanations, often at a slightly higher level than the main text. A reader not familiar with DFT is advised to skip most of the footnotes, but a reader familiar with it may find some of them useful ‘The paper is not meant to be a scholarly review of DF'T, but rather an informal guide to its conceptual basis and some recent developments and advances. 'The Hohonberg-Kohn theorem and the Kohn-Sham equations are discussed in some detail. Approximate density functionals, selected aspects of applications of DFT, and a variety of extensions of standard DFT are also discussed, albeit in less detail. Throughout it is attempted to provide ‘8 balanced treatment of aspects that are relevant for chemistry and aspects relevant for physics, bub with a strong bias towards conceptual foundations. ‘The texct is intended to be read before (or in parallel with) one of the many excellent more technical reviews available in the literature. ‘The author apol- ogives to all researchers whose work has not received proper consideration, ‘The limits of the author’s knowledge, as well as the limits of the available space and the nature of the intended audience, have from the outset prohib- ited any attempt at comprehensiveness.t TX fiat version of this text was published in 2002 as » chapter inthe proceedings of the VIII'th Summer School on Blectronie Structure of the Brastin Physical Society {i ‘The text-was unexpectedly wel received, and repeated requests from users prompted the author to electronically publish revised, updated and extended versions in the preprint archive hetp,//arav.org/arekive/eond-ma, where the second (2008), third (2004) and fourth (2005) versions were deposited under the reference number eond-mab/0211443. ‘The present ith (2006) version of this text, published inthe Brualian Journal of Phyoics, is approximately 60% longer than the first. Although during the consecutive revisions tmany embarrassing mistakes hve been removed, and unclear passages improved upon, ‘many otier doubtlewly remain, and much beautifal and important work has not ben ‘mentioned even in pasting, The return from electronic publishing to printed publishing, however, marie the completion ofa eye, and is intended to also mark the end of the 2 What is density-functional theory? Density-functional theory is one of the most popular and suecessful quan- tum mechanical approaches to matter. Tt is nowadays routinely applied for calculating, e.g., the binding energy of molecules in chemistry and the band structure of solids in physics, First applications relovant for fields tradition- ally considered more distant from quantum mechanics, such as biology and mineralogy are beginning to appear. Superconductivity, atoms in the focus of strong laser pulses, relativistic effects in heavy elements and in atomic nu- clei, classical liquids, and magnetic properties of alloys have all been studied with DFT. DFT owes this versatility to the generality of its fundamental concepts and the fiexibility one has in implementing them. In spite of this flexibility ‘and generality, DF'T is based on quite a rigid conceptual framework. ‘This, section introduces some aspects of this framework in general terms. The following two sections, 3 and 4, then deal in detail with two core elements of DFT, the Hohenberg-Kohn theorem and the Kohn-Sham equations. The final two sections, 5 and 6, contain a (necessarily less detailed) description of approximations typically made in practical DI*T calculations, and of some extensions and generalizations of DPT. ‘To get a frst idea of what density-functional theory is about, itis useful to take a stop back and recall some elementary quantum mechanies. In quantum mechanics we learn that all information we can possibly have about a given ystom is contained in the system’s wave function, W, Here we will exclusively be concerned with the olectronic structure of atoms, molecules and solids. ‘The nuclear degrees of freedom (e.g., the crystal lattice in a solid) appear only in the form of a potential u(r) acting on the electrons, so that the wave function depends only on the electronic coordinates.* Nonrelativistically, this wave function is calculated from Schridinger's equation, which for a single electron moving in a potential u(r) reads ny? [ oe +09] V(r) = W(t). aw If there is more than one electron (i.e., one has 2 many-body problem) ‘author's work on the Binds Bye View of Densily-Functional Theory, This is the so-called Born-Oppenhelmer approximation. Tt it common to call x(x) a ‘potential although i is, strictly speaking, a potential eneray. 4 Schrdinger’s equation becomes pe(- ESE a a, a) +n] Wr ra.. tw) = BW (rare. tN)s 7 SG @) where NV is the mumber of electrons and U(ri,1j) is the clectron-cleetron interaction. For a Coulomb system (the only type of system we consider here) one has PT Note that this is the same operator for any system of particles interacting via the Coulomb interaetion, just as the kinetic energy operator FFU) = Dw @) a ( Fs aa o is the same for any nonrelativistic system.* Whether our system is an atom, a molecule, or a solid thus depends only on the potential v(r,). For an atom, eB =D ap © Ire where Q is the nuclear chargot and R the nuclear position. When dealing with a single atom, R is usually taken to be the zero of the coordinate system. For a molecule or a solid one has Ln) where the sum on k extends over all nuclei in the system, each with charge Q = %e and position Ry. It is only the spatial arrangement of the Ry. (together with the corresponding boundary conditions) that distinguishes, fundamentally, a molecule from asolid.® Similarly, itis only through the term v ment ©) Irs for materials containing atoms with Inrge atomic number Z,, accelerating the electrons to relativistic velocities, one must include relativistic offects by solving Dirac’s equation fr an approximation to it. In this ease the kinetic energy operator takes a different form. ‘In Germs of the elementary charge e > O and the atomic number Z, the nuclear charge Js Q = Ze and the charge on the electron is S0ne sometimes says that P and Gare ‘universal, while 7 is system-dependent, ot “nonuniversal. We will come back to this terminology. 5 O that the (essentially simple) single-body quantum mechanics of Eq. (1) differs from the extremely complex many-bocly problem posed by Eq. (2). ‘These properties are built into DFT in a. very fandamental way. ‘The usual quantum-mechanical approach to Schrédinger’s equation (SE) can be summarized by the following sequence ovis fC) i.e., one specifies the system by choosing u(r), plugs it into Schrédinger's equation, solves that: equation for the wave function W, and then calculates observables by taking expectation values of operators with this wave function, ‘One among the observables that are calculated in this way is the particle density nan fdbra [dry... f Prm¥(era---stn) VO R---s1w). (8) Many powerful mothods for solving Schrisdinger’s equation have been devel- oped during decades of struggling with the many-body problem. In physics, for example, one hes diagrammatic perturbation theory (based on Feynman diagrams and Green’s functions), while in chemistry one often uses config- uration interaction (CI) methods, which are based on systematic expansion in Slater determinants. A host of more special techniques also exists. The problem with these methods is the great. demand they place on one’s compu- tational resources: it is simply impossible to apply them efficiently to large and complex systems. Nobody has ever calculated the chemical properties of a 100-atom molecule with full CI, or the electronic structure of a real semiconductor using nothing but. Green’s functions.® A simple estimate of the computational complexity of this task is bo imagine a reel space representation of ¥ on a mesh, in which each coordinate is discretized by using 2 mesh points (which is not very much). For N electrons, W becomes a function of 3. ‘coordinates (ignoring spin, and taking W to be real), and 20° values are required 20 ‘describe W on the mesh, The density n(r) is a fonction of three coordinates, and requires 26? values on the same mesh, GI and the Kohn-Sham formulation of DFT additionally ‘employ sets of single-particle orbitals. N such orbitals, used to build the density, require 20? N values on the same mesh. (A CI calculation employs also unoccupied orbitals, and roquites more values.) For N= 10 electrons, the many-body wave function thus requltes 20°°/20" = 10% times more storage space than the density, and 209/10 x 20%) m2 10% times more than sets of single-particle orbitals. Clovor uso of symmetries can reduce these ratiog, but the full many-body wave funetion rervains unaceassible for real systems with more than a fer electrons, v(t) 23 W(r, 72... tn) SS obsers It is here where DFT provides a viable alternative, less accurate por- haps,” but much more versatile, DIT explicitly recognizes that, nonrela- tivistic Coulomb systems differ only by their potential u(r), and supplies a prescription for dealing with the universal operators 1 and W once and for all.® Furthe:more, DPT provides a way to systematically map the many- body problem, with J, onto a single-body problem, without U. All this is done by promoting the particle density n(r) from just one among many ob- servables to the status of key variable, on which the calculation of all other observables can be based. ‘This approach forms the basis of the large ma- jority of eleotronic-structure calculations in physics and chemistry. Much of what we know about the electrical, magnetic, and structural properties of materials has been calculated using DIT, and the extent to which DFT has contributed to the science of molecules is reflected by the 1998 Nobel Prize in Chemistry, which was awarded to Walter Kohn (3), the founding father of DFT, and John Pople [4], who was instrumental in implementing DFT in ‘computational chemistry. ‘The density-functional approach ean be summarized by the sequence n(r) => W(n1,.-- tN) =e v(t), @ i.c., knowledge of n(r) implies knowledge of the wave function and the po- tential, and hence of all other observables, Although this sequence describes the conceptuel structure of DFT, it. does not really represent what is done in actual applications of it, which typically proceed along rather different lines, and do not make explicit use of many-body wave functions. The following chapters attempt to explain both the conceptual structure and some of the TAccuracy is a relative term, As a theory, DFT is formally exact. Ths performance in actual applications depends on the quality of the approximate density functionals ex- ployed. For small numbers of particles, or systems with special symmetries, essentially ‘exact solutions of Sehrédinger's equation ean be obtained, and no approximate functional ‘ean compete with exact solutions. For more realistic systems, modern (2005) sophisticated density fnictlonals attain rather high accurney. Data on atoms are colloctod in Table 1 in Soe. 6.2. Bonel-longths of molecules can be predicted with an average error of less than 0,00inm, lattice constants of solids with an average error of less than 0.005nm, and molec- ‘lar energies to within less than 0.2eV [2]. (For comparison: already a small molecule, such as water, has « total energy of 2081.1eV}. On the ober hand, energy gaps in solids cean be wrong, by 100%! ‘We will se that int practice this proseription can be implemented only approximately. Stil, these approximations retain a high degree of universality in the sense Ut they often ‘work well for mare than one type of system. many possible shapes and disguises under which this structure appears in applications. ‘The literature on DFTs large, and rich in excellent reviews and overviews. Some representative examples of full reviews and systematic collections of research papers are Refs. (5-19]. The present overview of DPT is much leas detailed and advanced than these treatments. Introductions to DFT that are more similar in spirit to the present one (but differ in emphasis and se- lection of topics) are the contribution of Levy in Ref. (9), the one of Kurth and Perdew in Refs. (15] and 16], and Ref. [20] by Makov and Argaman. My aim in the present text is to give a bird's-eye view of DFT in a language that should be accassible to an advanced undergraduate student: who has com- pleted a first course in quantum mechanics, in either chemistry or physics. Many interesting details, proofs of theorems, illustrative applications, and exciting developments had to be left out, just as any discussion of issues that are specific to only certain subfields of either physics or chemistry. All of this, and much more, can be found in the references cited above, to which the present little text may perhaps serve as a prelude. 3 DFT as a many-body theory 3.1 Functionals and their derivatives Before we discuss density-functional theory more carefully, let us introduce ‘usefull mathematical tool. Since according to the above sequence the wave function is determined by the density, we can write itas Y = U[nl(r,r2,...tw), which indicates that V is afunetion of its NV spatial variables, but-a functional of n(r). Functionals. More generally, a functional F[n] can be defined (in an admittedly mathematically sloppy way) as a rule for going from a function to a number, just as a function y = f(c) is a rule () for going from a number (2) to anumber (y). A simple example of a functional is the particle number, N= [érnte) = Nin), (10) which is a rule for obtaining the number N, given the function n(r). Note that the name given to the argument of n is completely irrelevant, since the 8 functional depends on the function itself, not on its variable, Hence we do not need to distinguish F(n(r)) from, e.g. F{n(r')]. Another important, case is that in which the functional depends on a parameter, such as in vl) = feet PED, a) which is a rule that for any value of the parameter r associates a value ‘uj|n](r) with the function n(r’). This term is the so-called Hartree potential, which we will repeatedly encounter below. Functional variation. Given a function of one variable, y = f(x), one can think of two types of variations of y, one associated with , the other with f. For a fixed functional dependence f(s), the ordinary differential dy measures how y changes as a result of a variation 2 — 2+ do of the variable «x. This is the variation studied in ordinary calculus. Similarly, for a fixed point 2, the functional variation 6y measures how the value y at this point changes as a result of a variation in the functional form f(z). ‘This is the variation studied in variational calculus. Functional derivative. The derivative formed in terms of the ordinary differential, df /dx, measures the first-order change of y = f(r) upon changes of 2, i.., the slope of the function f(z) at «: H(a+ de) = se) + Lae + Olas?) (12) The functional derivative measures, similarly, the first-order change in a func- tional upon a functional variation of its argument: PIC) +87(o)) = FU) + [ s(o)5fl@)de+ OC), (18) where the integral arises because the variation in the functional F is deter- mined by variations in the function at all points in space. ‘The first-order coefficient (or ‘functional slope’) (2) is defined to be the functional derivative FLA 5F(2), ‘The functional derivative allows us to study how a functional changes upon changes in the form of the function it depends on. Detailed rules for calculating functional derivatives are described in Appendix A of Ref. (6). A general expression for obtaining Functional derivatives with respect to n(x) of a functional F[n} = [ f(n,n!,n",n!,...;)dx, where primes indicate ordinary 9 derivatives of n(x) with respect to 2, is [6] cE [a] ea Oise cnO fa reano jae nO sar (a4) Gn(e) ~ On” da On! de® Onl ~ de® Bw * ‘This expression is frequently used in DFT to obtain xe potentials from xc energies.” The Hohenberg-Kohn theorem ‘At the heart of DFT is the Hohenberg-Kohn (HK) theorem, ‘This theo- rem states that for ground states Hq, (8) can be inverted: given a ground- state density ng(r) it is possible, in principle, to calculate the corresponding, ground-state wave function Vo(t1,r2..., rw). ‘This means that Wo is a func- tional of nq. Consequently, all ground-state observables are functionals of , too. Tf Yo can be calculated from nq and vice versa, both functions are equivalent and contain exactly the same information. At first sight this seems impossible: how can a function of one (vectorial) variable r be equivalent to a function of IV (vectorial) variables r...rw? How can one arbitrary variable contain the same information as N arbitrary variables? ‘The crucial fact which makes this possible is that knowledge of rto(t) implies implicit knowledge of much more than that of an arbitrary func- tion f(r). The ground-state wave function Yq must not only reproduce the ground-state density, but also minimize the energy. For a given ground-state density no(r), we can write this requirement as yo = prin (OP +0 +019), (15) where E,,o denotes the ground-state energy in potential u(r). The preceding ‘equation tells us that fora given den: state wave fungtion Wo is that which reproduces this Fan arbitrary density n(r), Bln] = pin(UIP +0 + V0). (16) "The use of functionals and their derivatives is not limited to density-functional ‘thoory, oF even to quantum mechanics. In classical mechanics, e.., one expresses the Lagrangian £ in terms of of generalized coordinates 9(2,t) and their temporal deriva- tives (zt), and obtains the equations of motion from extresnizing the action functional Alg] = f'£(a,4;8)dt. The resulting equations of motion are the well-known Ruler-Lagrange equations 96 — 96, which are a special ease of Bo (14) 10 If n is a density different from the ground-state density no in potential u(r), then the W that produce this n are different from the ground-state wave function ‘Vp, and according to the variational principle the minimum obtained from F,[n} is higher than (or equal to) the ground-state energy By = 2, [no]. ‘Thus, the functional B,(n} is minimized by the ground-state density no, and its value at the minimum is Ey. ‘The total-energy functional can be written as B,fn] = ming mn (DP + O[Y) + fd rn(r) F(nj+Vinj,| (17) where the intemal-energy functional F{n} = ming ..(W{P' + O|) is inde- pendent of the potential v(r), and thus determined only by the structure of the operators U und T. This universality of the internal-energy functional allows us to define the ground-state wave function Wo as that antisymmetric N-particle function that delivers the minimum of F[n[ and reproduces ‘9. e G0 ronclegenerate (for the case OF degeneracy soe foot- note 12}, this double requirement uniquely determines Wo in terms of no(r), without having to specify u(r) explicitly." Equations (15) to (17) constitute the constrained-search proof of the Hohenberg-Kohn theorem, given independently by M. Levy [22] and E. Lieb [23]. ‘The orjginal proof hy Hohenberg and Koln {24} proceeded by assuming that Vo was not determined uniquely by no and showed that this produced a contradiction to the variational principle. Both proofs, by constrained search ‘and by contradiction, are elegant and simple. In fact, it is a bit surprising that it took 38 years from Schrédinger’s first papers on quantum mechanics [25] to Holenberg and Kohn’s 1964 paper containing their famous theorem (24) ‘Since 1964, the HK theorem has been thoroughly scrutinized, and several alternative proofs have been found. One of these is the so-called ‘strong form of the Hohenberg-Kohn theorem’, based on the inequality [26, 27, 28] jj &rAn(r)Av(r) < 0. (aa) Here Av(r) is a change in the potential, and An(r) is the resulting change in the density, We see immediately that if Au #0 we cannot have An(r) = 0, Note that this is exnctly the oppasite of te conventional prescription to specify the Hamiltonian via v(2), and obtain Wo from solving Schr8dinger’s equation, without having to specify n(z) explicitly ul i.e., a change in the potential must also change the density. This observation implies again the HK theorem for a single density variable: there cannot be two local ith the same ground-state charge density. A given Noparticle ground-state density thus determines uniquely the corresponding potential, and hence also the wave function. Moreover, (18) establishes & relation betwoon the signs of An(r) and Av(r): if Av is mostly positive, An(r) must be mostly negative, so that their integral over all space is nega- tive. This additional information is not immediately available from the two classic proofs, and is the reason why this is called the ‘strong’ form of the HK theorem. Equation (18) can be obtained along the lines of the standard HK proof (26, 27], but it: can be turned into an independent. proof of the HK thoorem because it can also be derived perturbatively (see, e.g., section 10.10 of Ref, [28]).. Anothep aver ment is valid only for Coulomb potentials. It is based on Hato’s theorem, Which states (29, 30] that for such potentials the electron detsity has a cusp Al the position of the nuclei, where it satisfies Ne (19) Here Ry, denotes the positions of the nuclei, %, their atomic number, and ay = h?/me? is the Bohr radius, For a Coulomb system one can thus, in principle, read off all information necessary for completely specifying the Hamiltonian directly from examining the density distribution: the integral over n(r) yields NV, the total particle number, the position of the cusps of n(r) are the positions of the nuclei, Ry; and the derivative of n(r) at these positions yields % by means of Eq, (19). This is all one needs to specify the complete Hamiltonian of Eq. (2) (and thus implicitly all its eigenstates). In practice one almost never knows the density distribution sufficiently well to implement the search for the eusps and calculate the local derivatives. Still, Kato’s theorem provides a vivid illustration of how the density can indeed contain sufficient information to completely specify a nontrivial Hamilto- nian." For future reference we now provide a. commented summary of the content of the HK theorem. ‘This summary consists of four statements: Note that, unlike the fall Hohenberg-Kohn theorem, Kato's theorem does apply only to superpositions of Coulomb potentials, and can therefore not be applied directly to the effective Kohn-Shamn potential 2 (1) The nondegenerate ground-state (GS) wave function is a unique func- tional of the GS density:! Wot, Fa.) Pw) = Ulnolr))- (20) ‘This is the essence of the HK theorem. As a consequence, the GS expectation value of any observable O is a functional of rig(r), too: 05 = Oral = (Wl nol OI Uhr). (ay (2) Perhaps the most important observable is the GS energy. ‘This energy Bug = Exlro] = (¥lro]|H¥lr0]), (22) where H = P+ 0+ V, has the variational property"? Eelnio] < Foln'h, (23) where no is GS density in potential V and n’ is some other density. This ig very similar to the usual variational principle for wave functions. From calculation of the expectation value of a Hamiltonian with a trial wave function " that is not its GS wave function Ve one can never obtain an enorgy below the tre GS energy, (Wolff ¥o) < (WEY) = BW (24) Eyo = Elo) Similarly, in exact DFT, if Pfn] for fixed v2: is evaluated for a density that is not the GS density of the system in potential vee, one never finds a result below the true GS energy. ‘This is what Ha, (23) says, and it is so impor- tant for practical applications of DF'P that if i sometimes called the second Hohenberg-Kohn theorem (Eq. (21) is the first one, then) 1R{f the ground state is degenerate, several of the degenorate groune-state wave functions ray produce the same density, 20 that unique fuxetional ln) does not exist, but by Aafinition these wave functions all yield Uhe same energy, so that the functional yf] continues to exist and to be minimized by ng. A universal functional Fn] eam alo sill be defined (5) 53h minimum of Bn is Uhos attained for the ground-state density. All other extrema ofthis functional comrespond to densities of excited slates, but the excited states obtained In this way do not nevessatily cover the entire spectrum of the many-body Hamiltonian (51, 13 In performing the minimization of B,{n] the constraint that the total par- ticle number NV is an integer is taken into account by means of a Lagrange multiplier, replacing the constrained minimization of E,[n] by an uncon- strained one of B,[n] — aN. Since N = f drn(r), this leads to Buln} OE. n(x) ON' (28) where js is the chemical potential. (3) Recalling thet the kinetic and interaction energies of a nonrelativistic Coulomb system are described by universal operators, we can also write By as Bul fn] + Ul] + Vin] = Pin) + Vn, (26) where T[n] and U[n] are universal functionals (defined as expectation values of the type (21) of P and O), independent of u(r). On the other hand, the potential energy in a given potential u(r) is the expectation value of Ea. (6), Vin] = J @rn(rju(r), (7) and obviously nonuniversal (it depends on u(r), ie., on the system under study), but very simple: once the system is specified, i.c., »(r) is known, the functional V[nj is known explicitly. (4) There is a fourth substatement to the HK theorem, which shows that if v(r) is not hold fixed, the functional V[n] becomes universal: the GS density determines not only the GS wave function Vo, but, up to an additive constant, also the potential V = V[nol. ‘This is simply proven by writing Schrédinger’s equation as H= Daley = 6 - Te, @s) which shows that any eigenstate W, (and thus in particular the ground state Wo = W{no]) determines the potential operator V up to an additive constant, the corresponding eigenenergy. As a consequence, the explicit reference to the potential v in the energy functional B,[n] is not necessary, and one can rewrite the ground-state energy as {re} Ina) (29) ‘Another consequence is that no now does determine not only the GS wave function but the complete Hamiltonian (the operators T and U are fixed), ‘and thus all excited states, too: We(ta,ra---)tw) = Vilna), (30) where & labels the entire spectrum of the many-body Hamiltonian 17. 3.3 Complications: N and v-representability of densi- ties, and nonuniqueness of potentials Originally the fourth statement was considered to be as sound as the other three, However, it has become clear very rovently, as a consequence of work of H. Bechrig and W, Pickett [32] and, independently, of the author with G. Vi- gnale (33, 34), that there are significant, exceptions to it. In fact, the fourth substatement holds only when one formulates DFT exclusively in terms of the charge density, as ve have done up to this point. It does not hold when one works with spin densities (spin-DFT) or current dénsities (current-DF'T).™ In these (and some other) cases the densities still determing the wave finc- tion, but they do not uniquely determine the corresponding potentials. This so-called nonuniqueness problem has been discovered only recently, and its consequences are now beginning to be explored (27, 32, 33, 34, 35, 36, 37, 38} It is clear, however, that the fourth substatement is, from a practical point of view, the least important of the four, and most applications of DFT do not have to be reconsidered as a consequence of its eventual failure. (But, some do: see Refs. (33, 34} for examples.) ‘Another conceptual problem with the HIC theorem, much better known and more studied than nonuniqueness, is representability. ‘To understand whet representability is about, consider the following two questions: (i) How does one know, given an arbitrary function m(r), that this function can be represented in the form (8), i.c., that it is a density arising from an antisym- metric N-body wave function (ry ....rj)? (ii) How does one know, given & finction that can be written in the form (8), that this density is a ground- state density of o.local potential v(r)? ‘The first of these questions is known as the W-epreemapiity problem, the second is called v-representability. Note that these are quite important questions: if one should find, for example, in a Win Section 6 we will briefly discuss hese formulations of DFT, numerical calculation, a minimum of By{n) that is not N-representable, then this minimum is not the be phys acceptable solution to the problem at, hand, Luckily, the single-particle density aati ‘No similarly genéfal Solution is known 7 the v-reprcsental (The HK theorem only guarantors that there connote mare than-ane-Do- tential for each density, but does not exclude the possibility that there is less than one, ie, 7270, po' | that in discretized systems every density is ensemble v-representable, which means that a local potential with a degenerate ground state can always be found, such that the density n(r) can be written as linear combination of the densities arising from each of the degenerate ground states [41, 42, 43}. It is not known if one of the two restrictions (‘discretized systems’, and ‘ensemble’) can be relaxed in general (yielding ‘in continuum systems’ and ‘pure-state’ respectively), but it is known that one may not relax both: there are densities in continuum systems that, are not representable by a single non- degenerate antisymmetric function that is ground state of a local potential u(r) (5, 41, 42, 43). In any case, the constrained search algorithm of Levy and Lieb shows that v-representability in an interacting system is not required for the proof of the HK theorem. For the related question of simultaneous: v-representability in a noninteracting system, which appears in the context of the Kohn-Sham formulation of DFT, see footnotes 34 and 35. 3.4 A preview of practical DFT After these abstract: considerations let us now consider one way in which one can make practical uso of DFT, Assume we have specified our system (i.e, u(r) is known). Assume further that we have reliable approximations for U{n] and T[n]. In principle, all one has to do then is to minimize the sum of kinetic, interaction and potential energies prt __» Bln) = Tho) + Ula} + Vin] = Tin) + Ula) + farneyo(e) (31) rae oo with respect to n{r). ‘The minimizing fimetion n9(r) is the system’s GS charge density and the value Kyo = Ey[nq] is the GS energy. Assume now that v(r) depends on a parameter a. This can bo, for example, the lattice constant in a solid or the angle between two atoms in a molecule. Calculation 16 ae of By for many values of a allows one to plot the curve Byo(a) and to find the value of a that minimizes it. This value, do, is the GS lattice constant or angle. In this way one can calculate quantities like molecular geometries and sizes, lattice constants, unit cell volumes, charge distributions, total energies, ete. By looking at the change of F%,o(a) with a one can, moreover, calculate compressibilities, phonon spectra and bulk moduli (in solids) and vibrational frequencies (in molecules). By comparing the total energy of a composite system (e.g., a molecule) with that of its constituent systems (e.g., individual atoms) one obtains dissociation energies. By calculating the total energy for systems with one olectron more or less one obtains electron affinities ‘and ionization energies." By appealing to the Hellman-Feynman theorem fone can calculate forces on atoms from the derivative of the total energy with respect to the nuclear coordinates. All this follows from DFT without having to solve the many-body Schridinger equation and without having to make a single-particle approximation, For brief comments on the most useful additional possibility to also calculate single-particle band structures see Secs. 4.2 and 4.2.3. In theory it should be possible to calculate all observables, since the HK theorem guarantees that they are all functionals of ng(r). In practice, one does not know how to do this explicitly. Another problem is that the minimization of B,[n) ig, in general, a tough numerical problem on its own And, moreover, one needs reliable approximations for T'{n| and Un] to begin with. In the next section, on the Kohn-Sham equations, we will see one widely used method for solving these problems. Before looking at that, however, it is worthwhile to recall an older, but still occasionally useful, alternative: the Lins Fas Pronto la Epproximation one sets: Viel =Ualn)=& [er fer EAD, (32) i.e., approximates the full interaction energy by the Hartree energy, the elec- trostatic interaction energy of the charge distribution n(r). One further Wilectron alinities are typically harder to oblain than ionlaation energies, because within the loeal-densty and generalized-gradient approximations the N+ I’st electron is too weakly bound or even unbound: the asymptotic effective potential obtained from these approximations decays exponentially, and not as 1/r, Le, it approaches zero 60 fast that binding of negative ions is strongly suppressed, Self-interaction corrections or other fully nonlocal functionals are needed to improve this behaviour 17 Ti approximates, initially, Th] as THAI} farvomnte)), (33) where er is the kinetic-energy density of a, homogeneous, interacting system with (constant) density yi, Since it Téfers to interacting electrons f(a) is not known explicitly, and further approximations are called for. As it stands, however, this formula is already a first example of a local- density approximation (LDA). In this type of approximation one imagines the real inhomogeneous system (with density n(r) in potential v(r)) to be decomposed in small cells in each of which n(r) and v(r) are approximately constant. In each cell (i., locally) one can then use the per-volume energy of a homogeneous system to approximate the contribution of the cell to the real inhomogeneous one. Making the cells infinitesimally small and summing over all of them yields Bq. (33). Fora. ified by subscript 5, for ‘single particle’) = RPP Ciom) irther approximate Tn] TIP An) we TEP] = if @rter(n(r)), (34) where T/?4in} is the local-density approximation to T;{n], the kinetic energy of noninteracting electrons of density n. Equivalently, it may be considered the noninteracting version of T“?4[n). (The quantity T,[n] will reappear below, in discussing the Kohn-Sham equations.) The Thomas-Fermi,approx- imation’® then consists. in combining (32) and (34) ‘and writing Tin] + Un) + Vin] © Bf) tt) +Unln) + Vb A major defect of the Thomas-Fermi approximation is that within it molecules are unstable: the energy of a sot of isolated atoms is lower than that; of the bound molecule. ‘This fundamental deficiency, and the lack of accuracy result- ing from neglect of correlations in (32) and from using the local approxima- tion (34) for the kinetic energy, limit the practical use of the ‘Thomas-Fermi (35) 187he Thomas- Fermi appreximation for screening, discussed in many books on solid- state physics, is obtained by minimizing B?{n} with respect to m and linearizing the resulting relation between v(x) and n(r). It thus involves one more approximation. (the linearization) compared to what is called the ‘Thomas-Fermi approximation in DFT [44]. In two dimessions no linearization is required and both become equivalent (44) 18 approximation in its own right. However, it is found a most useful starting point for a large body of work on improved approximations in chemistry and physics [12, 30]. More recent approximations for T'jn] can be found, e.g, in Refs. [45, 46, 47]; Tit The Context of orbital-Iree DFT. "The extensioi of the local acisity ConeoHT ur the exchange-corTelation Ghergy is at the heart of many modern density funetionals (see Sec. 5.1). 3.5 From wave functions to density functionals via Green’s functions and density matrices It is a fimdamental postulate of quantum mechanics that the wave function contains all possible information about a system in a pure state at zero tem- perature, whereas at nonzero temperature this information is contained in the density matrix of quantum statistical mechanics. Normally, this is much more information that one can handle: for a system with N = 100 parti- cles the many-body wave function is an extremely complicated function of 300 spatial and 100 spin!” variables that would be impossible to manipulate algebraically or to extract any information from, even if it were possible to calculate it in the first place, For this reeson one searches for less complicated objects to formulate the theory. Such objects should contain the experimen- tally relevant information, such as energies, densities, etc., but do not need to contain explicit information about the coordinates of every single parti- cle. One class of such objects are Green’s functions, which are described in the next subsection, and another are reduced density matrices, described in the subsection 3.5.2. Their relation to the wave function and the density is summarized in Fig. 1. 3.5.1 Greon’s functions Readers with no prior knowledge of (or no interest in) Green’s functions should skip this subsection, which is not necessary for understanding the following sections, In mathematics one usually defines the Green’s function of a linear op- erator £ via [z — L(r)|G(a,2/;2) = 62 — 2), where 6(x — a’) is Dirac's TTP keep the notation simple, spin labels are either ignored or condensed into a common (rs) in most of this text. They will only be put back explicitly in discussing, /-funetional theory, in Sec. 6 19 delta function, For a single quantum particle in potential u(r) one has, for example, = le + ca vt GOV(r, 5s B) = hd(r —r') (36) Many applications of such single-particle Green's functions are discussed in Ref. [21]. In many-body physics it is nseful to also introduce more compli- cated Green’s functions. In an interacting system the single-particle Green’s function is modified by the presence of the interaction between the particles.™* In general it, now satisfies the equation’? a wy s [ag +5 veo} (6, tyr! t!) = hale — eat — t+) “i i, Bx U(r — x)GO"l (rt, x; r't’, xt), (37) where G® (rt, xt; r’t’, x¢*) is the two-particle Green’s function (21, 48]. Only for noninteracting systems (U = 0) is G(r,t;r’,) a Green’s funetion in the mathematical sense of the word, In terms of G(r,4;1/,t/) one can explicitly express the expectation value of any single-body operator (such as the po- tential, the kinetic energy, the particle density, ete.), and also that of certain two-particle operators, such as the Hamiltonian in the presence of particle- particle interactions. One way to obtain the single-particle Green's function is via solution of what is called Dyson’s equation [21, 48, 49), Gl, r,t) = G(r, Gr',2) + ii ae f aa! / Br i &r! GO, 5x, 7) N06, 7, G75 0,0), (38) where © is known as the irreducible self energy (21, 48, 49] and G® is the Green's function in the absence of any interaction. This equation (which ‘Note that expressions like “wo particle operator and ‘single-particle Green's function? refer tothe numberof particles involved inthe defiition ofthe operator (two in the ease of an interaction, one fora potential energy, et), not to the total number of partes present in the aya, "When energy i conserved, ie, tha Hamiltonian does not depend om time, Gt!) depends on time only via the difference {—1' and can be writtan as (1:0). By Foutier teansformation with respect lo t~ one then passes rom G(r3L-— 1) to Clays) of Eq, (35). ee arbitrary two-particle operators one needs the full two-particle Green’s function oo 20 ‘we will not attempt to solve here) has # characteristic property that we will meet again when we study the (much simpler) Kohn-Sham and Hartroo-Fock equations, in Sec. 4: the integral on the right-hand side, which determines G on the left-hand side, depends on G itself, The mathematical problem posed by this equation is thus nonlinear, We will return to such nonlineatity when we discuss self-consistent solution of the Kohn-Sham equation. ‘The quantity 3 will appear again in Sec. 4.2.3 when we discuss the meaning of the eigenvalues of the Kohn-Sham equation, ‘The single-particle Green's function is related to the irreducible self en- ergy by Dyson's equation (38) and to the two-particle Green’s function by the equation of motion (37). It can also be related to the ze potential of DFT by the Sham-Schliiter equation [50] [aroce) J wG (0, 1w)G(r, rj) = [eet [bet wcuce, rw) ecle' ri) O(ue" r5), (39) where G, is the Green's function of noninteracting particles with density n(x) (i., the Green’s function of the Kohn-Sham equation, see See. 4.2), and Dye(r’, 1”; w) = D(r’, 2”; w) — 5(r — r)uyr(r’) represents all contributions to the full irreducible self energy beyond the Hartree potential, ‘A proper discussion of 3 and G requires formalism known as second quantization (21, 48] and usually proceeds via introduction of Feynman dia- grams, ‘These developments are beyond the scope of the present overview. A related concept, density matrices, on the other hand, can be discussed easily. ‘The next section is devoted to a brief description of some important density matrices. 3.5.2 Density matrices For a general quantum system at temperature 7’, the density operator in a canonical ensemble is defined as exp Pf Trlexp-PA)’ where Tr['] is the trace and = 1/(kp‘7’). Standard textbooks on statistical physics show how this operator is obtained in other ensembles, and how it is used to calculate thermal and quantum expectation values, Here we focus (40) aL on the relation to density-functional theory. To this end we write ¥ in the energy representation as — Lv0xp7P™ Wi) al es (41) where |;) is eigenfunction of #7, and the sum is over the entire spectrum of the system, each state being weighted by its Boltzmann weight exp", At zexo temperature only the ground-state contributes to the sums, so that F= Wah (42) ‘The coordinate-space matrix element of this operator for an N-particle sys- tem is (215 Bay tw Flths Bay tty) = W(t, Ba, Bw) Y(t, 22, ty) Ss yey ay 5 hy May ty) (43) which shows the connection between the density matrix and the wave funo- tion, (We use the usual abbreviation x = rs for space and spin coordi- nates.) The expectation value of a general N-particle operator O is obtained from O = (0) = fda; J dea... f day U(x, 22, .0w)"OW (ar, m2, 20m), which for multiplicative operators becomes (O)= [des [doa f day Oy(2r,2,.2ni21.0002n) (44) and involves only the function 9(st1,2, 25 1522,-0N), which is the di- agonal element of the matrix y. Most operators we encounter in quantum mechanics are one or bwo-partiele operators and can be calculated from re- duced density matzices, that depend on less than 2N variables?" The reduced two-particle density matrix is defined as sole ene 8) = N(N = 1 MONEY fds f daa f denen, 26 02,24 34,245.) 18) Ws as or Gren non express Ie vo parti operat and oo patil dovty mesic sor tosh mab af pars ned inte tatnin tUw oper (too in the case of an interaction, one for a potential energy, et), not to the total number ‘of particles prasent in the system, 22 where IV(V ~ 1)/2 is a convenient normalization factor. This density ma- trix determines the expectation value of the particle-particle interaction, of static correlation and response functions, of the xe hole, and some re- lated quantities. The pair-correlation function g(x, 2’), eg., is obtained from the diagonal element of ya(c1, #2; 2%;,4%,) according to qo(a1y 52,2 ni(xs)n(2a)o(22"). Similarly, the single-particle density matrix is defined as N f din [ aes | ae WN f dae f day f dea. [ den ¥" (22,255 80)Y( 24, 20,2355) (46) aor ti) = | dewey, 5 245.2052, 22,4524) 2H) ‘The structure of reduced density matrices is quite simple: all coordinates that +7 does not depend upon are set equal in W and W*, and integrated over. The single-particle density matrix can also be considered the time-independent form of the single-particle Green's funetion, since it can alternatively be obtained from a2!) = —ilim G(x, 2! ~ (47) In the special case that the wave function W is a Slater determinant, ive., the wave function of N noninteracting fermions, the single-particle density matrix can be written in terms of the orbitals comprising the determinant, a a.) = Dagy(oosle’) (4) 7 which is known as the Dirac (or Dirac-Fock) density matrix. ‘The usefulness of the single-particle density matrix becomes apparent. when we consider how one would calculate the expectation value of a mul- tiplicative single-particle operator A = SY a(r;) (such as the potential V = EN o(e)): w CA) [t6,.-[ don "(ener ven) [Sale] Mean v2n) 49) [ar f dew V(e4, 20, 8m a{ar)Wr,22n) (50) [eole)re,z), (61) 23, which is a special case of Ba. (44). ‘The second line follows from the first by exploiting that the fermionic wave function W changes sign upon interchange of two of its arguments. The last equation implies that if one knows +(2,2) one can calculate the expectation value of any multiplicative single-particle operator in terms of it, regardless of the mumber of particles present in the system. The simplification is enormous, and reduced density matrices are very popular in, e.g., computational chemistry for precisely this reason. More details are given in, o.g,, Ref. [6]. The full density operator, Faq. (40), on the other hand, is the central quantity of quantum statistical mechanics It is not possible to express expectation values of two-particle operators, such as the interaction itself, or the full Hamiltonian (i.e., the total energy), explicitly in terms of the single-particle density matrix y(r,1’). For this purpose one requires the two-particle density matrix. This situation is to be contrasted with that of the single-particle Green’s function, for which one nows how to express the expectation values of ( and Hf (21, 48]. Apparently, some information has gotten lost in passing from G toy. This can also be seen. very clearly from Eq. (47), which shows that information on the dynamics of the system, which is contained in G, is erased in the definition of 4(r, r). Explicit information on the static properties of the system is contained in the N-particle density matrix, but ai seen from (45) and (46), a large part of this information is also lost (“integrated out’) in passing from the N-particle density matrix to the reduced two- or one-particle density matrices. ‘Apparently even less information is contained in the particle density” n(r), which is obtained by summing the diagonal element of (2, a’) over the spin variable, ale) = Des, 28) (2) ‘This equation follows immediately from comparing (8) with (46). We can define an altcrnative density operator, fi, by requiring that the same equation must also be obtained by substituting fi(r) into Bq, (51), which holds for any single-particle operator. This requirement implies that fi(r) = DN 6(r—r,).2* for nonmmltipicativo singlo-partiele operators (such as the kinetic energy, which con- tains a derivative) one requires the full single-particle matrix (¢,/) and not only (x, 2). 14 quantitative estimate of how much less information is apparently contained in the density thon in the wave fanction is given in footnote 6, “The expectation value of is the particle density, and therefore Ais often also called the density operator. This coneopt must not be confused with any of the various density ‘matrices or the density operator of statistical physics, Ea. (40) 24 ‘The particle density is an even simpler function than y(x,2’): it de- pends on one set of coordinates « only, it can easily be visualized as a three- dimensional charge distribution, and it is directly accessible in experiments. ‘These advantages, however, seem to be more than compensated by the fact that one has integrated out an enormous amount of specific information about the system in going from wave functions to Green’s functions, and on to density matrices and finally the density itself, This process is illustrated in Fig. 1. ‘The great surprise of density-functional theory is that in fact no informa- tion has been lost at all, at least as long as one considers the system only in its ground state: according to the Hohenberg-Kohn theorem the ground- state density rio(z) completely determines the ground-state wave function Wola, a2...) 25 Hence, in the ground state, a function of one variable is equivalent to a function of NV variables! ‘This property shows that we have only integrated out: explicit information on our way from wave functions via Green’s functions and density matrices to densities. Implicitly all the information that was contained in the ground-state wave function is still contained in the ground- state density. Part of the art of practical DFT is how to get this implicit information out, once one has obtained the density! 4 DFT as an effective single-body theory: The Kohn-Sham equations Density-functional theory can be implemented in many ways. ‘The min- imization of an explicit. energy functional, discussed up to this point, is not normally the most efficient among them. Much more widely used is the Kohn: Sham approach. Interestingly, this approach owes its success and popularity partly to the fact that it does not exclusively work in téfinis of thepurticle (0 charge) density, but brings a special kind of wave functions (single-particle orbitals) back into the game. As a consequence DFT then looks formally like a single-particle theory, although many-body effects are still included The Runge Gross theorem, which forms the basis of time-dependent DP [61], sim arly guarantees Uhat the time-dependent density contains the same information as the time-dependent wave function 3 oe, | oo meat | Yeon fn YF Getty) Figure 1: Information on the timo-and-space dependent wave func- tion U(x1,52...,20y,t) is built into Green’s functions, and on the time- independent. wave function into density matrices. Integrating out degrees of freedom reduces the N-particle Green's function and N-particle density matrix to the one-particle quantities G(t;,2;t) and (11,22) described in the main text. The diagonal element of the one-particle density matrix is the ordinary charge density — the central quantity in DFT. The Hohenberg- Kohn theorem and its time-dependent: generalization (the Runge-Gross the- orem) guarantee that the densities contain exactly the same information as the wave functions, via the so-called exchange-correlation functional. We will now see in some detail how this is done. 4.1 Exchange-correlation energy: definition, interpre- tation and exact properties 4.1.1 Exchange-correlation energy ‘The Thomes-Fermi approximation (34) for '{n] is not very good. A more accurate scheme for treating the kinetic-energy functional of interacting elec- trons, Tal, is based on decomposing it into one part that represents the ki- notic energy of noninteracting particles of density n, e., the quantity called above T,[n], afd One that represents the remainder, denoted Ty[n} (the sub- presents tt eiconated| 26 scripts » and ¢ stand for ‘single-particle’ and ‘correlation’, respectively). oe To) = Tale] + Tele (53) T,[n) is not known exactly as a functional of m [and the LDA to ap- proximate it leads one back to the Thomas-Fermi approximation (34)], but it is easily oxpressed in terms of the single-particle orbitals 4,(r) of a nonin- teracting system with d icc Se sea Tip because for noninteracting particles the total kinetic energy is just the sum. of the individual kinetic energies. Since all ¢,(r) are functionals of n, this expression for T, is an explicit orbital functional but an implicit density functional, T,(n] = 7,[{ in|}, where the notation indicates that T, depends ‘on the full sel of occupied orbitals ¢, each of which is a functional of n. Other such orbital functionals will be discussed in Sec. 5. ‘We now rewrite the exact energy functional as Bln] = Tho] + Vln] + Vir) = Talila) s] + Ursin] + Pacln| + Vind) (55) where by definition #,. contains the differences T —T; (ie. T.) and U — Uy. ‘This definition shows that@ significant part ofthe correlation en Ee is due to the difference 7, bétwean the noninteracting ant the interacting, Kinetic energiés: Unlike’ Eq-(35), Eq. (55) is formally exact, but of course Fe is unknown — although the HK theorem guarantees that it is a density functional. This functional, Eyc{n], is called the exchange-correlation (xc) energy. It ig often decomposed a8 Hye = By + Be, where Hy is due to the Pauli principle (exchange energy) and 2, is due to correlations. (: is then a part of Z,,) The exchange energy can be written explicitly in terms of a) Je-particlo orbitals ax 7 THOR A OAM CL (56) ——& | Bel{diln}) 7, i deined ag the expectation value of the kinetioensray operator 7 with the Slater determinant avising from density nm, Le., Tsfa) = (O{n)[Z|®[a)). Similarly, the full Iinetic energy is defined as Tt] = (Wfa)f7}¥[u). All consequences of antiymmetrization (i.e., exchange) are described by employing a determinantal wave function in defining 7. nce, Tn, the difference between T; and T is a pure correlation effct. "This differs from the exchange energy used in Hartree-Fock theory only in the substi- tution of Hartree Fock orbitals 4f!"() hy Kohn-Sham orbitals #(). 27 which is known as the Fock term, but no general exact expression in terms of the density is known a 4.1.2 Different perspectives on the correlation energy For the correlation energy no general explicit expression js known, neither in terms of orbital nor densities. Different waya to understand correlations are described below. Correlation energy: variational approach, A simple way to understand the origin of correlation is to recall that the Hartree energy is obtained in ‘2 variational calculation in which the many-body wave function is approx- imated as e product of single-particle orbitals, Use of an antisymmetrized product (2 Slater determinant) produces the Hartree and the exchange energy [48, 49]. ‘The correlation energy is then defined as the difference between the full ground-Stave ehergy [obtatred with the correct, many-body wave func- tion) and the one obtained from the (Harteeo-Fock or Kohn-Sham®) Slater determinant. Since it arises from a more general trial wave function than a single Siater determinant, correlation cannot raise the total energy, and Ein] <0. Since a Slater doterminant is iiself more general than a simple product we also have Bz <0, and thus the upper bound” B,,{n} < 0. Correlation energy: probabilistic approach. Recalling the quantum me- chanical interpretation of the wave function as a probability amplitude, we see that a product form of the many-body wave function corresponds to treat- ing the probability amplitude of the many-clectron system as a product of the probability amplitudes of individual electrons (the orbitals). Mathemat- ically, the probability of a composed event is only equal to the probability of the individual events if the individual events are independent (i.e., uncor- related). Physically, this means that the electrons described by the product wave function are independent." Such wave functions thus neglect the fact that, as a consequence of the Coulomb interaction, the electrons try to avoid ™The Hartree-Fock and the Kohn-Sham Slater determinants are not identical, since they ate compoeed of different single-particle oral, and thus the definition of exchange and correlation energy in DFT and in conventional quantum chemistry i slightly diffrent 159) 24 lower bound is provided by the Lieb-Oxford formula, given in Eq. (64). Correlation is a mathematical concept describing the fact that certain events are not indegerent="tt can be defined also in classical physics, and in applications of statistics to other fields than physics. Exchange is due to the indistinguishability of particles, and ‘8 true quantum phenomenoff, without any analogue in classical physics. 28 Se! each other. The correlation energy is simply the additional energy lowering obtained in a real system due to the mutual avoidance of the interacting electrons. One way to characterize a strongly correlated system is to define correlations as strong when B, is comparable in magnitude to, or larger than, other energy contributions, such as Byy or T,. (In weakly correlated systems , normally is several orders of magnitude smaller.) Correlation energy: beyond mean-field approach. A rather different (but equivalent) way to understand correlation is to consider the following alter- native form of the operator representing the Coulomb interaction, equivalent to Ba, (3), Dak fete far HOO) AOE), ol r= rf in which the operator character is carried only by the density operators ft (in occupation number representation), and the term with the delta function subtracts out the interaction of a charge with itself (¢f., e.g., the appendix of Ref. (53) for a simple derivation of this form of J). The expectation value of this operator, U = (W|O|¥), involves the expectation value of a product of density operators, (Wli(r)i(r’)/W). In the Hartree term (32), on the other hand, this expectation value of a product is replaced by a product of expectation values, each of the form n(r) ~ (WlA(r)|W). This replacement amounts to a mean-field approximation, which neglects quantum: fluctuations about the expectation values: by writing # = n+ dfigte and substituting in Eq, (57) we see that the difference between (||) and the Hartree term (32) is entirely due to the fluctuations dfiyiye and the self- interaction correction to the Hartree term. Quantum fluctuations about the expectation value are thus at the origin of quantum correlations between interacting particles, ‘conduction band in a solid with the kinetic energy (if tf band widtlpissmaller, correlations ‘are strong), or the quasiparticle energies & with the Koln-Bhatii eigenvalues ¢ (if both are similar, correlations are weak, see footnote 37), or the derivative discontinuity Ane, fined in Hq, (66), with the Kohn-Sham energy gap (if the former is eomparable to or lnrgor than the letter, correlations are strong). (‘The meaning of @, ¢ and Ba is explained below.) No universally applicable definition of ‘strong correlations? seems to exist. At finite temperature there are also thermal fluctuations. ‘Tb properly include these fone must use & fnite-tomperature formulation of DPT (54). See also the contribution of B. L. Gyorffy otal, in Ref. 19] for DFT treatment of various types of fluctuations Other characterizations of strongly correlated wr compare the width of the 29 Correlation energy: holes. ‘The fact that both exchange and correlation tend to keep electrons apart, has given rise to the concept of an ze hole, ge(t, 1), describing the reduction of probability for encountering an lec- tron atv, given one at r, ‘The ac energy can be written as a Hartree-like interaction between the charge distribution n(r) and the #¢ hole tge(t,’) = tha(t, 7) + Melts"), Balnh © fete fate Mee) (68) r-rel] which defines nize. The exchange component E,[n] of the exact exchango- correlation functional describes the energy lowering due to antisymmetriza- tion (i.e, the tendency of like-spin electrons to avoid each other). It gives rise to the exchange hole ng(r, x’), which obeys the sum rule fdr! tg(r,r!) = —1 ‘The correlation component H,{n] accounts for the additional energy lower- ing arising because electrons with opposite spins also avoid each other. The resulting correlation hol egrates to zero, so that the total ac hole satis- fies f Pr! rge(r,r!) = —1. The ze hole can also be written a8 tye(r, 1) n(e")(Gln|(x, x) — 1), where g is the average of the pait-correlation function g(r,r), mentioned in Sec. 3.5.2, over all values of the particle-particle interac- tion, from zero (KS system) to (1) (interacting system). This average is sim- ply expressed in terms of the coupling constant a as 9(r,1’) = o ga(r,”}da. For the Coulomb interaction, e?, ie., the square of the electron charge (5, 6) 4,13 Exact properties Clearly E; is an enormously complex object, and DFT would be of little use if one had to know it exactly for making calculations, The practical advantage of writing £{n] in the form Eq. (65) is that the unknown functional Baefn| is pigally much smaller than the known terms Ts, Uj and V. One can thus Choos) ‘at reasonably simple approximations for Fi,<{n] provide useful results ‘Some successful approximations are discussed in Sec. 5. Exact: prop- erties, such as the sum rule [ d°rn,.(r,r') = —1, described in the preceding section, are most valuable guides in the construction of approximations to Bye Among the known properties of this functional are the coordinate scaling conditions first obtained by Levy and Perdew [55] (59) 30 Ex{ny] > Bdnj| ford > 1 (60) Bn] < ABe{n] ford <1, (61) where f(r) = \°n(n) is a scaled density integrating to total particle number N. Another important property of the exact functional is the one-electron limit, ~ aS. Ban] = 0 . ©) Fl) = Filo), (63) where n) ig a one-electron density. ~“These-tatter-two-cenditions, which are satisfied within the Hartree-Fock approximation, but not by standard lgcal- density and gradient-dependent functionals, ensurd that there 1s no artificial SIEIaLSraDhon oF Ore Sco WEES = ‘The Lieb-Oxford bound [56, 57], Eg{n] > Ezeln] > —1.68¢? f Prax), (64) establishes a lower bound on the xc energy, and is satisfied by LDA and many (but not all) GGAs. One of the most intriguing properties of the exact functional, which has resisted all attempts of describing it in local or semilocal approximations, is the derivative discontinuity of the ec functional with respect to the total particle number (50, 58, 59], Saar "| 6Bsln] a(n) | = v(t} — z(t) = Ave \ (68) wes Ont) Ins where 6 is an infinitesimal shift of the electron number N, and Age is a systom-dependent, but r-independent shift of the ae potential tec(r) as it passes from the electron-poor to the electron-rich side of integer N. The ing kinetic-energy functional has a similar discontinuity, given by where ey and ey41 are the Kohn-Sham (KS) single-particle energies of the highest occupied and lowest unoccupied eigenstate. ‘The meaning of these al KS eigenvalues is discussed in the paragraphs following Ha. (75) and illus trated in Fig. 2. In the chemistry literature these are called the HOMO (highest occupied molecular orbital) and LUMO (lowest unoccupied molec- ular orbital), respectively. ‘The kinetic-energy discontinuity is thus sim- ply the KS single-particle gap Axs, or HOMO-LUMO gap, whereas the ac discontinuity Age is a many-body effect. ‘The true fundamental gap A = B(N +1)-+E(N — 1) —2E(N) is the discontinuity of the total ground- state energy functional (5, 50, 58, 59], A= el Since all terms in £ other than E,, and T, are continuous functionals of n(r), the fundamental gap is the sum of the KS gap and the ac discontinuity. Sjandard density Functionals (LDA and GGA) predict Aye = 0, and thus Shen inaeresenmats the fundamental gap ~The tomcementat ind KS gaps are also illustrated in Fig. 2. All these properties serve as constraints or guides in the construction of approximations for the functionals ,[n] and B.(n]. Many other similar prop- erties are known. A useful overview of scaling properties is the contribution of M, Lovy in Ref. [19]. $n] las ~ _ = Ast Boe (7) 4.2. Kohn-Sham equations 4: 1 Derivation of the Kohn-Sham equations Since T; is now written as an orbital fimctional one cannot directly minimize Bq. (55) with respect to n. Instead, one commonly employs a scheme sug- gested by Kohn and Sham (60] for performing the minimization indirectly. ‘This scheme starts by writing the minimization as SU pln] , 5Bscln) _ 8TaIn} a Fale) Tamla) * ule) * Sale) dn(ey CP IC) Healt) xo (68) ‘As a consequence of Eq. (27), 6V/in “GEE eet potenti the electrons move in.® ‘The term 6Uy/én siniply Vise Wie Hartree potential, = s| 3| SSThie potential is called ‘external’ because it is external to the electron system and not generated solf-consistently from the clectron-eloctron interaction, as vr aud Uys. Ib comprises the lattice potential and any additional truly external field applied to the system a a whole 32 introduced in Bg. (11). Por the term 5¥ye/5n, which can only be calculated explicitly once an approximation for Hye has been chosen, one commonly ‘writes Uae. By means of the Sham-Sehliiter equation (39), tye is related to the irreducible solf energy ©, introduced in Eq, (38) [60] Consider now a system of noninteracting particles moving in the potential s4(t). Por this systom the minimization condition simply ——~- Bain] 6Tafr] | SVe[n] _ ST[n] Salt) ~ Sn(e) * Inks) ~ ney 7 (09) since there are no Hartree and xc terms in the absence of interactions. The density solving this Euler equation is n(r). Comparing this with Eq. (68) wwe find that both minimizations have the same solution n4(r) = (1), if v ig chosen to be = < TES [tale = ve) + ont) tect). (70) Consequently, one can Zaloulate. the. density. of-the Tiiteracting (many-body) System in potential v(r), described by a many-body Schrodinger equation of the form (2), by solvi equations of a noninteracting (single-body) system in potential v,{r)." — In particular, the Schrddinger equation of this auxiliary system, HE" + vale)] dilt) = edule), x (7) yields orbitals that reproduce the density n(r) of the original system (these Sore tine sume orbitals employod IEG. (54)), EI AiO, (72) where fi is the occupation of the ith orbital." Eqs. (70) to (72) are the celebrated Kohn-Sham (KS) equations. ‘They replace the problem of minj- mizing E[n] by that of solving a noninteracting fiiger equation. (Recall ‘The question whether such a potential vg(r) always exists bv tho mathomadical suse is called the noniuterneting v-ropresentabilty problem. It is known that every interacting ‘ensemble v-representable density is also noninteracting ensemble v-representable, but, as ‘mentioned in See. 3.2, only in discretived rystoms has it been proven that all densities axe interacting ensemble e-ropresentable. It is not known i interacting ensemble-representabe dnsities may be noninteracting, pute-state representable (ie, by « single determinant), ‘which would be convenient (but is not neesssary) for Kohn-Sham calculations. Normally, the occupation numbers f; follows an Aufaw principle (Fermi statistics) with f= 1 lori M, and 0S fy $1 fori = N (ie, at most n(r) 33. that the minimization of Hn) originally replaced the problem of solving tho many-body Schrédinger equation!) Since both vjr and Yge depend on n, which depends on the gy, which in turn depend on v4, the problem of solving the KS equations is a nonlinear cone, just as is the one of solving the (much more complicated) Dyson equation (38). The usual way of solving such problems is to start with an initial guess for n(r), calculate the corresponding v,(r), and then solve the differential equation (71) for the gy. From these one calculates a now density, using (72), and starts again, The process is repeated until it converges. ‘The technical name for this procedure is ‘self-consistency cycle’. Different convergence criteria (such as convergence in the energy, the density, or some observable calculated from these) and various convergence-accelerating algorithms (such as mixing of old and new effective potentials) are in common use. Only rarely it requires more than a few dozen iterations to achieve convergence, and even rarer are cases where convergence seems unattainable, ie., a self-consistent solution of the KS equation cannot be found. ‘Once one has converged solution no, one can calculate the total energy from Eq. (55) or, equivalently and more conveniently, from®® Bo = DY = F fbr f de! PALEY — Far vee(r)no(t) + Beelnol.| (73) Equation (73) follows from writing V(n] in (55) by means of (70) as [bre (eyn(e) = fbr fuse) —vnle) —vele}ia(e) (74) = Vain) ~ far fon e)-+ vele)Io(e), (75) Vb the highest occupied orbital can have fractional occupation). Some densities that are not noninteracting wrepresentable by a single ground-state Slater determinant, may still be ‘obtained from a single determinant if one uass occupation numbers Je that leave holes below the HOMO (the Fermi energy in a metal), so that J: # 1 even for some i < NV {31}, but this is not guaranteed to daseribe all possible densities. Alternatively (soe Soe. 8.2 and footnote 84) a Kohn-Sham equation may be sct up in terms of ensembles of determinants, ‘This guarantees noninteracting v-representabilty forall densities that azo interacting ensemble v-representable. For practical KS calculations, the most important, consequence ofthe fact that not every arbitrary density is guaranteed to be noninterecting, ‘representable is that the Kohn-Sham selfeonsistency cycle is no’ guaranteed to converge. PAI terms on the right-hand side of (73) except for the first involving the sum of the single-particle energies, are sometimes known os double-counting corrections, in analogy to a similar equation valid within Hartree-Fock theory. 34 and identifying the energy of the nonintoracting (Kohn-Sham) system as B= Dla =Tet Ve 4.2.2. The eigenvalues of the Kohn-Sham equation Equation (73) shows that Hy is not simply the sum* of all ¢. In fact, it should be clear from our derivation of Ba, (71) that the ¢ are introduced as completely artificial objects: they are the eigenvalues of an auxiliary single- body equation whose eigenfunctions (orbitals) yield the correct density. Tt is only this donsity that has strict physical meaning in the KS equations. The KS eigenvalues, on the other hand, in general bear only a semiquantitative resemblance with the true energy spectrum [61], but are not to be trusted quantitatively ‘The main exception to this rule is the highest occupied KS eigenvalue. Denoting by ev(M) the Nt eigenvnnte Of a system with electrons, one can show rigorously that ey(N) = =/, the negative of the first ionization energy of the N-body nti wr the negative of the electron affinity of the iy sysiem 158, 02, 63]. ‘These relations hold for the gxtotTynctional only When calculated with an approximate functions? of the LDA or GGA type, the highest eigenvalues usually do not provide good approximations to the experimental J and A. Better results for these observables are obtained by calculating them as total-energy differ- ences, according to I = By(N — 1) ~ Bo(N) and A = Ey(N) — Ba(N +1), where Ep(N) is the ground-state energy of the N-body system. Alterna- tively, sel-interaction corrections can be used to obtain improved ionization energies and electron affinities from Kohn-Sham eigenvalues [64] Figure 2 illustrates the role played by the highest occupied and lowest umoceupied KS eigenvalues, and their relation to observables. For molecules, HOMO(N) is the highest-occupied molecular orbital of the N-olectron sys tem, HOMO(N+1) that of the N + L-elecizon system, and LUMO(N) the lowest unoccupied orbital of the N-electron systerh. In solids with a gap, 7The difference between Ky and: 7" ¢ is due to particle-particle interactions. The ad- dion ternw on tho right-hand ide of (73) gve mathematical meaning tothe cominon statment th Ue whole o more Un the su oe pars It By ca be wien approx: Imatly aw 51" @ (whore the aro not ihe sume mi the KS eigenen ) the eta cane denetbed in tors of Mealy inverecting uaspertichs, euch wh eneray ce Fermbiguid tory in mata and eflectivemas teary i somicondictors tre exemple ofthis type of approach Figure 2: Schomatic description of some important Kohn-Sham eigenvalues relative to the vacuum level, denoted by 0, and their relation to observables, ‘See main text for explanations. the HOMO and LUMO become the top of the valence band and the bottom. of the conduction band, respectively, whereas in metals they are both iden- tical to the Fermi level. The vertical lines indicate the Kohn-Shamn (single: particle) gap Ax, the fundamental (many-body) gap A, the derivative dis- continuity Of thé xc functional, Ay, the ionization enexgy of the interacting Neeleetron system I(N) = —ey(N) (which is also the ionization energy of the Kohn-Sham system Jics(IN)), the electron affinity of the interacting NV- electron system A(N) = —ew (NW +1) and the Kohn-Sharn electron affinity Axs(N) = ~ensi WV). Given the auxiliary nature of the other Kohn-Sham eigenvalues, it comes as a groat (apd welcome) surprise that in many situations (typically char- acterized by the presence of fermionic quasiparticles and absence of strong correlations) the Kohn-Sham eigenvalues ¢ do, empirically, provide a rea- sonable first approximation to the actual energy levels of extended systems. ‘This approximation is behind most band-structure calculations in solid-state physics, and often gives results that agree well with experimental photoemis- sion and inverse photoemission data [65], but much research remains to be done before it is clear to what extent such conclusions can be generalized, and how situations in which the KS eigenvalues are good starting points for approximating the true excitation spectrum are to be characterized micro- 36 scopically (66, 67]. Most band-structure calculations in solid-state physics are actually caleu- lations of the KS eigenvalues e. This simplification has proved enormously successful, but when one uses it one must be aware of the fact that one is taking the auxiliary single-body equation (71) literally as an approximation to the many-body Schrédinger equation. DFT, practiced in this mode, is not a rigorous many-body theory anymore, but a mean-field theory (albeit one with a very sophisticated mean field vs(r)). ‘The energy gap obtained in such band-structure calculations is the one} called HOMO-LUMO gap in molecular calculations, i.e., the difference be- tween the energies of the highest occupied and the lowest unoccupied single- particle states. Neglect of the derivative discontinuity Age, defined in Fa. (65), by standard local and semilocal 2), on the other hand, are exceedingly difficult to calculate, and little is known about them. In this situation it was a major breakthrough when it, was realized, in the early eighties, that instead “Of power-series-ke systematic gradient expan- sions one could experiment with more general functions of n(r) and Vn(r), which need not proceed order by order. Such functionals, of the general form Bao _ J er f(n{r), Vato), (ss) have become Imown as genevalized-gradient, approximations (GGAs) [95] Different GGAs differ in the choice of the function J(m, Vn). Note that this makes different GGAs much more different from each other than the different parametrizations of the LDA: essentially there is only one correct “Sif one adds this term to the ‘Thomas-Fermi expression (36) one obtains the s0-callod “Thomas-Pormi-Welasickor approximation to Bj. In a systematic gradient expansion th 8 in the denominator is roplacad by a 72 [5 6). “Remarkably, the form of this term is fully determined already by dimensional analysis! In BSEA®) — ¢? Fr F(r4|VnP) the function f must have dimensions (length)~*. Since ‘the dimensions of n and |Vn[? are (length)~* and (length)~®, respectively, and to second ‘order no higher powers oF higher derivatives of n are allowed, the only possible combination is fo [n(e)F/nt expression for el(n), and the various parametxizations of the LDA (89, 90, 92, 93, 94) are merely different ways of writing it. On the other hand, depending on the method of construction employed for obtaining f(n, Vin) ‘one can obtain very different GGAs. Ih particular, GGAs used in quantum chemistry typically proceed by fitting parameleis 16" teat"wIs"OF Elected moleauTes” Ui The oUNer” Windy CCAS tse fi RYE Cem "W"EMPhasize REE aals NowaaapT the nos popi a a a TOTS] CAs are PBE (denoting the functional proposed “if T9896 BY ‘Pardew, Burke and tha che iy ts physies are Bey (aencting tis com fe GombinNTION OF Weekes 18 eatOT ‘ils available, = now ones 6 Quite generally, current GGAs seem to give reliable results for all main types of chemical bonds (covalent, ionic, metallic and hydrogen bridge), For, yan der Waals interactions, however, ,.common. GGAs and LDA fail. To de- sci RC TRESFAELIOHY several more specialized approaches have been developed within DFT (101, 102, 103, 104, 105]. Both in physics and in chemistry the widespread use of GGAs has lend to major improvements as compared to LDA. ‘Chemical accuracy’, as defined above, has not yet been attained, but is not too far away cither. A nseful collection of explicit expres- sions for some GGAs can be found in the apps dix of | of Ref: Ji06], and more dSERNOT TRACT Of Some selected CCAR Gad their performance is given in Ref. [107] and in the chapter of Kurth and Perdew in Refi. (15, 16) No systematic attempt at comparing explicit functionals can be made here, but many detailed comparisons are available in the literature. For pure illustrative purposes only, Table 1 contains ground-state energies of the ‘Ar atom, obtained with several of the methods discussed previously in this chapter. Footnote 7 contains additional information on the performance of DFT for larger systems. the PBB GGA [96] and the TPSS MGGA (2] (s0e below) may be partial exeeptions (09, 100) because they work reasonably well near the equilibrium distance of the van der ‘Waals bond, but they recover only the short-range behnnviour and do not deseribe correctly the loug-raige asymptotic regime of the van der Waals interaction. aT method -E/ou. Thomas-Fermi | 625.7 Hartree-Fock 526.818 OEP (exchange only) | 526.812 LDA (exchange only) | 524.517 LDA (VWN) 525,946 LDA (PW92) 525,940 LDA-SIG(P2) 528,993 ADA 527,922 WDA 528.957 GGA (B88LYP) | 527.551 “experiment 527.6 ‘Table 1: Ground-state energy in atomic units (1 au. = 1 Hartree = 2 Rydberg = 27.21eV ~627.Skcal/mol) of the Ar atom (Z = 18), obtained with some representative density functionals and related methods. The Hartree- Fock and OEP(exchange only) values are from Krieger et al. (third of Ref. {120)), ADA and WDA values are from Gunnarsson et al., Ref. [129], as reported in Ref. [5], and the LDA-SIC(PZ) value is from Perdew and Zunger, Ref, [93]. ‘The experimental value is based on Veillard and Clementi, J. Chem. Phys. 49, 2415 (1968), and given to less significant. digits than the calculated values, because of relativistic and quantum electrodynamical effects (Lamb shift) that are automatically included in the experimental result, but not in the calculated values 48, 5.3. Orbital functionals and other nonlocal approxima- tions: hybrids, Meta-GGA, SIC, OEP, ete. In spite of thase advances, the quest for more accurate functionals goes ever on, and both in chemistry and physies various beyond-GGA functionals have appeared. Perhaps the most popular functional in quantum chemistry is BSLYP. This is combination of the LY! lation [98] with e-Foek exchange into 5 possible). ‘The con- struction of hybrid functional involves certain amount of empiricism in the choice of fimetionals that are mixed and in the optimization of the weight actors given to the HF and DFT terms. Formally, this might be considered a drawback, but in practice B3 has proven to be the most successful exchange functional for chemical applications, in particular when combined with the LYP GGA functional for B,, More extreme examples of this semiempirical mode of construction of functionals are Becke’s 1997 hybrid functional (109), which contains 10 adjustable parameters, and the functionals of Refs. [110] and [111], each of which contains 21 parameters ‘Another recent bevond-GGA development is the emergence of so-called Mela-GGAs, which depend, in addition to the density and its derivatives, ‘also on the Kohn-Sham kinetic-energy density 7(r) [2, 112, 113] x m Mineticencrey Comey. a(t) = 5 Vee), (89) so that By, can be written as Hyc(n(r), Vn(r),r(0)}. ‘The additional degree of frecdom provided by 1 is usd to satisfy additional constraints on Fc, such as a selFinteraction-corrected correlation functional, recovery of the fourth-order gradient expansion for exchange in the limit of slowly varying, densities, and a finite exchange potential at the nucleus [2]. In several recent tests [2, 100, 114, 115, 116] Meta-GGAs have given favorable results, even when compared to the best GGAs, but the full potential of this type of approximation is only beginning to be explored systematically. ‘AS we have seen in the case of T,, it can be much easier to represent a functional in terms of single-particle orbitals than directly in terms of the FLPhis was weitten in early 2002, but ab Uhe time of revision of this text in 2006 it is sll correct. 49 density. Such functionals are known as orbital functionals, and Eq, (54) con- stitutes a simple example. Another important orbTarTpendent, functional the exchange energy (Fock term) of Eq. (56). The Meta-GGAs and hybrid functionals mentioned above are also orbital functionals, because they de- pend on the kinetic energy density (89), and on « combination of the orbital functional (56) with ordinary GGAs, respectively. Atill_another type of orbital functional is the self-interaction correction (SIC). Most implementations of SIC take tse of tie expressions-proposed in Ref. (93) (PZ-SIC), Berens tn, ny) which subtracts, orbital by orbital, the contribution the Hartree and the xe fiineTfonals would make if there was only one electron in the system. This correction can be applied on top of any approximate density finctional, and ensures that the resulting corrected functional satisfies Zg7re"S!C[y(), 0] = —Byfn®) for a one-clectron system. ‘The LDA is exact for a completely uniform system, and thus is selE-interaction free in this limit, but neither it nor common GGAs satisfy the requirement of freedom from self-interaction in general, and even Meta-GGAs have a remaining self-interaction error in their exchange part [2, 112]. This gelfinteraction is particularly critical for localized states, such as the d states in transition-metal oxides. For such systems PZ-SIC has been shown to greatly improve the uncorrected LDA {71, 72], but for thermochemistry PZ-SIC does not seem to be significant {117} Unfortunately the PZ-SIC approach, which minimizes the corrected en- ergy Fiction WR FXpEN Ue Srbitals, does not lead to Kohn-Sham equations of the usual form, because the resulting effective potential is dif ferent for each orbital. As a Gondéqiiénce, Various specialized algortthtay for nithimizing~the-PZ:SIC energy functional have been developed. For more details on these algorithms and some interesting applications in solid-state physics soe Refs. [71, 72, 73]. For finite systems, PZ-SIC has also been im- plemented by moans of the OBP {64, 74], which produces a common local potential for all orbitals, and is discussed in the next paragraph. A detailed review of implementations and applications of PZ-SIC can be found in the contribution of Temmerman ct al. in Ref, [17]. Alternatives to the PZ-SIC formulation of Ref, {93} have recently been analysed in [118, 119), with a view on either improving results obtained with PZ-SIC, or simplifying the Paring, my] — 3) (Bilin Fxg" |nig,0]), (90) 50 implementation of the correction. Since hybrid functionals, Meta-GGAs, SIC, the Fock term and all other orbital functionals depend on the density only implicitly, via the orbitals dn), it is not possible to directly calculate the functional devivative voc = 6B,0/6n, Instead one must use indirect approaches to minimize E[n} and obtain vze. In the case of the kinetic-energy functional T,[{¢ifn]}] this in- direct approach is simply the Kohn-Sham scheme, described in See. 4. In the case of orbital expressions for Bye the corresponding indirect scheme is known as the optimized effective potential (OEP),[120} or, equivalently, the optimized-potential model (OPM). (121). "The minimization of the orbital functional with respect to the density is achioved by repeated application of the chain rule for functional de: srselto}) _ ~én(r) SEC] Bul) Bee") Shr) bu(r") Sn(r) Yeeln|(r) = (91) where 22? is the orbital functional (eg., the Fock term) and v, the KS effective potential, Further evaluation of Iiq. (91) gives rise to an integral equation that determines the v,¢[n] belonging to the chosen orbital func- tional Bee{{¢i{n)}] (120, 122]. As an alternative to solving the full OEP integral equation, Krieger, Li and Iafrate (KLI) have proposed a simple but surprisingly accurate approximation that greatly facilitates implementation of the OEP (120]. ‘The application of the OBP methodology to the Fock term (56), either with or without the KLI approximation, is also known as the exact-exchange method (EXX). The OBP-EXX equations have been solved for atoms (120, 121, 123] and solids (124, 125), with very encouraging results. Other orbital- dependent functionals that have been treated within the OEP scheme are the PZ self-interaction correction (64, 74] and the Colle-Salvetti functional [123]. A detailed review of the OEP and its KLI approximation is Ref. [122] ‘The high accuracy attained by complex orbital functionals implemented via the OEP, and the fact that it is easier to devise orbital functionals than explicit density functionals, makes the OEP concept attractive, but the com: putational cost of solving the OEP integral equation is a major drawback. However, this computational cost is significantly reduced by the KLI approx imation [120] and other recently proposed simplifications (126, 127, 128]. In the context of the EXX method (ie., using the Fook exchange term as orbital functional) the OEP is a viable way to proceed. For more complex orbital 51 functionals, additional simplifications may be necessary (120, 126, 127, 128). A further reduction of computational comploxity is achieved by not eval- uating the orbital functional self-consistently, via Eq. (91), but only once, using the orbitals and densities of a. converged self-consistent LDA or GGA calculation, This ‘post-GGA’ or ‘post-LDA’ strategy completely avoids the OEP and has been used both for hybrid functionals and Meta-CiGAs (108, 109, 112, 113]. A drawback of post methods is that they provide only ap- proximations to the selfeonsistent total energies, not to eigenvalues, effective potentials, orbitals or densities. In the case of hybrid functionals, still another mode of implementation has become popular. ‘This alternative, which also avoids solution of Pq. (91), is to calculate the derivative of the hybrid functional with respect to the single- particle orbitals, and not with respect to the density as in (91), The resulting single-particle equation is of Hartree-Fock form, with a nonlocal potential, and with a weight factor in front of the Fock term. Strictly speaking, the orbital derivative is not what the HK theorem demands, but rather a Hartree- Fock like procedure, but in practice it is a convenient and successful approach. ‘This scheme, in which self-consistency is obtained with respect to the single- particle orbitals, can be considered an evolution of the Hartree-Fock Kohn- Sham method [6], and is how hybrids are commonly implemented. Recently, it, has also been used for Meta-GGAs [2]. For occupied orbitals, results obtained from orbital selfconsistency differ little from those obtained from the OEP. Apart from orbital functionals, which are implicit nonlocal density func- tionals because the orbitals depend on the density in a nonlocal way, there is also a class of explicit nonlocal density functionals. Such nonlocal density functionals take into account, at any point r, not only the density at that point, n(r), and its derivatives, Vn(r) etc., but also the behaviour of the density at different points r’ # r, by means of integration over physically relevant regions of space. A typical example is BAP An] = far n(e)eter(n(e), (92) where ci?” is the per-particle wc energy of the homogeneous electron liquid (see footnote 47). In the LDA one would have fi(r) = n(r), but in the average-density approximation (ADA) one takes (129] a(n) = jj br! neu (Ir —2')), (93) 52 where wln]([r — r')) is a weight fimetion that samples the density not only semilocally, as do the GGAs, but over a volume determined by the range of w. Conceptually similar to the ADA is the weightec-density approximation (WDA) (129). In terms of the pair-correlation function (see Secs. 3.5.2 and 4.1.2) the LDA, ADA and WDA functionals ean be written as 2 A [ate fev ene ion bata |(e £ a arf pct Gromlii(e)(e—)~1) (95) & far eve Crom{ate)ir—e)-1), (96) 1) (94) where in each case Jhom(r — r’) is the pair-correlation function of the homo- geneous electron liquid, averaged over the coupling constant ¢ (5, 6). ‘The dependence of these functionals on f(r), the integral over n(r), in- stead of on derivatives, as in the GGAs, is the reeson why such functionals are called nonlocal. In practice, this integral turns the functionals compu- tationally expensive, and in spite of their great promise they are much less used than GGAs. However, recent comparisons of ADA and WDA with LDA and GGAs for low-dimensional systems [114, 130] and for bulk silicon [131] show that nonlocal integral-dependent density functionals can outperform local and semilocal approximations. 6 Extensions of DFT: New frontiers and old problems Up to this point we have discussed DFT in terms of the change (or par- ticle) density n(r) as fundamental variable. In order to reproduce the cor- rect charge density of the interacting system in the noninteracting (Kohn- Sham) system, one must apply to the latter the effective KS potential v. — + vy + te, in which the last two terms simulate the effect of the clectron- electron interaction on the charge density. ‘This form of DFT, which is the one proposed originally (24), could also be called ‘chaxge-only’ DFT. It is not the most widely used DFT in practical applications, Much more common is a formulation that employs one density for each spin, my(r) and n(x), 53. ie, works with two fundamental variables. In order to reproduce both these in the noninteracting system one must now apply two effective pote tials, vp(t) and v,(r)." "This formulation of DFT is known as spin-DFT (SDFT) (89, 90]. Tts fundamental variables mj(r) and nj(r) can be used to ealeulate the charge density n(r) and the spin-magnetization density m(r) from n(r) = my(r) + my(r) (97) m(r) = wolny(r) — n(x), (98) where jz = gh/2me is the Bohr magneton, More generally, the Hohenberg- Kohn theorem of SDFT states that in the presence of a magnetic field B(r) that couples only to the electron spin {via the familiar Zeeman term Jdrm(r)B(z)] the ground-state wave function and all ground-state obsery- fables are unique functionals of m and m or, equivalently, of my and ny. Almost the entire further development of the HK theorem and the KS equa- tions can be immediately rephrased for SDF, just by attaching a suitable spin index to the donsities. For this reason we could afford the Iuxury of exclusively discussing ‘charge-only’ DFT in the preceding sections, without missing any essential aspects of SDFT. ‘There are, however, some exceptions to this simple rule. One is the fourth statement of the HK theorom, as discussed in Sec. 3.2. Another is the construction of functionals. For the exchange energy it is known, e.g,, that (#2! (204) + B2"T am) (99) fe Mey to the coordinate scaling of Bigs. (69) - (61), this property is of ten called ‘spin-scaling’, and it can be used to construct an SDFT exchange functional from a. given DFT exchange funetional. In the context of the ISDA, von Barth and Hedin [89] wrote the exchange functional in terms of an interpolation between the unpolarized and fully polarized electron gas which by construction satisfies Eq, (99). Alternative interpolation procedures "More generally, one requires one effective potential for each density-like quantity to be reproduced in the KS system. Such potentials and corresponding densities are called ‘conjugate variables ‘Bp the pacticular case B = 0 the SDFT HK theorem sill holds and continues to b uweful, ef, for systems with spontanoous polarization. In principle one could also use) ‘charge-only’ DPT to study auch systems, but then (2) and m,(r) hocome functionals o ‘n(r) and nobody knows how to determine these functionals, 54 or SDF. A generalization of DFT that does account for spin-orbit coupling and other relativistic effects is relativistic DFT (RDFT) (137, 138]. Here the fundamental variable is the relativistic four-component current ji", RDFT requires @ more drastic reformulation of DFT than does SDF. In particu- lar, the KS equation of RDFT is now of the form of the single-particle Dirac ‘equation, instead of the Schrédinger equation. ‘There are also many subtle ‘questions involving renormalizability and the use of the variational princi- ple in the presence of negative energy states. For details on these problems and their eventual solution the reader is referred to the chapters by Engel ct al, in Refs, [10] and [19], and to the book by Eschrig (18). A didactical exposition of RDFT, together with representative applications in atomic and condensoc-matter physics, can be found in the book by Strange [28], and a recent numerical implementation is presented in Ref. [139]. ‘To study the magnetic properties of matter one would often like to be able to obtain information on the currents in the system and their coupling to pos- sible external magnetic fields. Important classes of experiments for which this, information is relevant are nuclear magnetic resonance and the quantum Hall effects, SDF'T does not provide explicit information on the currents. RDFT. in principle does, but standard implementations of it are formulated in a spin-only version, which prohibits extraction of information on the currents. Furthermore, the formalism of RDFT is considerably more complicated than that of SDFT. In this situation the formulation of nonrelativistic current- DFT (CDFT), accomplished by Vignale and Rasolt [140, 141], was ¢ major step forward. CDFT is formulated explicitly in terms of the (spin) density and the nonrelativistic paramagnetic current density vector jp(r). Some re- cont applications of ODFT are Refi. (142, 148, 144, 145]. E, K, U, Gross and the author have shown that the existence of spin currents implies the exis- tence of a link between the «c functionals of SDFT and those of CDPT (146) Conceptually, this link is similar to the one of Eq, (99) between functionals of DFT and SDFT, but the details are quite different, Some approximations for zc functionals of CDT are discussed in Refs, [146, 147, 148). In addition to SDFT, RDFT and CDFT, there exist many other gene:- alizations of DFT that were designed for one or other special purpose. As ‘examples we mention superconductivity (149, 150, 151; 152] and spin-density ‘waves [136, 153], bub there are many more [5-19]. For reasons of space we ‘cannot discuss these extensions here, Instead, Jet us take a brief look ab a problem that roquires more radical departures from the framework of con- ventional DET: excited states, DFT is formulated in terms of ground-state 56 can be found in Ref. {92]. GGA exchange functionals also satisfy Eq. (99) by construction. For the correlation energy no scaling relation of the type (99) holds, so that in Tn practice correlation fuietionnts-are-eitherdiveatly con structed in terms of thé Spin densities or written by using, ‘Without formal justifientiony the Stine Interpolation already used in the exchange Tanetional in the case of the LSDA this latter procedure was introduced in Ref. [89], and further analysed and improved in Ref. {92]. ‘The Kohn-Sham equations of SDFT are 2 [EE ol] fel) = eet, (200) where Ygq(t) = v(t) + v(t) + Ysee(t). In a nonrelativistic calculation the Hartree term does not depend on the spin label,‘ but in the presence of ‘an externally applied magnetic field v,(r) = u(r) — oj40B (where o = +1). Finally, SESE Tiny m4} Sno(r) In the presence of an internal magnetic field Bye (i.e., in spin-polarized sys- toms) Yao, ~ ser ~ HoBze. This field is the origin of, e.g., ferromagnetism in transition metals. References to recent work with SDF'T include almost all practical DFT. ation: SDIT is by far the most widely wed form of DFT. Soine recent work on SDFT is described in Ref. [133]. A more detailed discussion of SDFT can be found in Refs. (5, 6, 90], and a partieu- larly clear exposition of the construction of zc functionals for SDFT is the contribution of Kurth and Perdew in Refs. (15, 16}. If the direction of the spins is not uniform in space one requires a for- mulation of SDFT in which the spin magnetization is not a scalar, as above, but a throo-component vector m(r). Different proposals for extending SDF'T to this situation are available (134, 135, 136). One mechanism that can give rise to noncollinear magnetism is spin-orbit coupling. This is another rela- tivistic effect (28), and as such it is not consistently treated in either DFT (01) Ueeo(P) = Spin-apin dipolar intoraction# are a relativistic effect of order (Ie), as are eurrent- current interactions. SSSDET has become synonymous with DFT to auch an extent that often no distinctio®] / fs made between the two, Le., a calculation referred to as a, DFT one is most of the time | really an SDFT one. wv 8Such ‘noncollinenr magnetism’ appanrs, 6.6, a8 canted! or helial spin configuestions in rare-earth compounds, or as domain walls in ferromagnets 55 functional theory can be applied to such model Hamiltonians, too, once a suitable density-like quantity has been identified as basic variable, Following, pioneering work by Gunnarsson and Schénhammer 170], LDA-type approx- imations have, o.g., recently been formulated and exploited for the Hubbard [171), the delta-interaction [172] and the Heisenberg [173] models. Common aspects and potential uses of DIT for model Hamiltonians are described in (74 Still another way of using DFT, which does not depend directly on ap- proximate solution of Kohn-Sham equations, is the quantification and clari- fication of traditional chemical concepts, such as electronegativity [6], hard- ness, softness, Fukui functions, and other reactivity indices [6, 175], or aro- maticity (176). ‘The true potential of DFT for this kind of investigation is only beginning to be explored, but holds much promise, ‘All extensions of DFT face the same formal questions (¢.g,, simultaneous interacting and noninteracting v-representability of the densities, nonunique- ness of the KS potentials, meaning of the KS eigenvalues) and practical problems (¢.g., how to efficiently solve the KS equations, how to construct accurate approximations to By., how to treat systems with very strong cor- relations) as do the more widely used formulations ‘charge-only’ DPT and SDF. These questions and problems, however, have never stopped DFT from advancing, and at present DFT emerges as the method of choice for solving a wide variety of quantum mechanical problems in chemistry and physics — and in many situations, such as large and inhomogeneous sys- tems, it is the only applicable first-principles method at all. ‘The future of DFT is bright: [3, 61, 177] — but to be able to contribute toil, the reader must now leave the present superficial overview behind, and turn to the more advanced treatments available in the literature (5-19) Acknowledgments The author has learned density-functional theory from E. K. U, Gross, and then practiced it in collaborations with B. L. Gyérify, L. N. Oliveira, and G. Vignale. ‘These scholars are in no way responsible for the content of this work, but the author's intellectual debt to them is enormous. Useful comments by J. Quintanilla, H. J. P. Freire, ‘T, Mar- casso, E, Orestes, N, A. Lima, N. Argaman, V. L. Libero, V. V. Franca, M. Odashima, J. M. Morbee, A. P. Favaro, A. J. R. da Silva and L. N. Oliveira on earlier versions of this manuscript are gratefully acknowledged. This work was supported financially by FAPESP and CNPq. 58 densities, and it_js not immediately obvious how one could extract informa- tion on excited states from them (althougir tr Teast in the case Oretmrperotily” DET the Tourth substatement.of the HK tigoren guarantees that this must be possible). ~Kpart from the ad hoc identification of the KS eigenvalues true exci- tation energies, thre cxists 4 considerable vartety of more-sound-approwches to excited states in DFT that have met with some success. ‘The early sug- gestion of Gunnarsson and Lundqvist [90] to use a symmetry-dependent ce functional to calculate the lowest-energy excited state of acl symmetry tins “tag Been implemented approximately by von Barth {154}, but suffers from. lack of knowledge on the symmetry dependence of the functional. More recent worl on this dependence is Ref. [155]. An alternative approach to excited states, not restricted to the lowest energy state of a given symmetry, is ensemble DFT, developed by Theophilou [26] and further elaborated by Oliveira, Gross and Kohn [156]. In this formalism the functional depends on. the particular choice for the ensemble, and a simple approximation for this dependence is available [156]. Some applications of this method have been worked out by Nagy [157]. Other DFT approaches to excited states can be found in Refs. {158}, [159], [160] and [31], but the most widely used method today is time-dependent DFT (TD-DFT). T dependent gencralization of the HK Theorem, the Runge-Gross theorem, cannot be proven along the lines of the original HK theorem, but requires a different approach (51, 161]. For recent reviews of ‘TD-DFT see Ref. [162]. Excited states have first been extracted from TD- DET in Refs. (163, 164]. This approach is now implemented in standard quantum-chemical DFT program packages [165, 166] and is increasingly ap- plied also in solid-state physies [70]. Another important application of TD- DFT is to aystems in external time-dependent fields, such as atoms in strong laser fields (167, 168]. First steps towards studying dynamical magnetic phe- nomena with TD-SDFT have been taken very recently {169}, All these extensions of DFT to time-dependent, magnetic, relativistic and a multitude of other situations involve more complicated Hamiltonians than the basic ab initio many-electron Hamiltonian defined by Eqs. (2) to (6) Instead of attempting to achieve a more complete description of the many- body system under study by adding additional terms to the Hamiltonian, it is often advantageous to employ the opposite strategy, and reduce the complexity of the ab initio Hamiltonian by replacing it by simpler models, which focalize on specific aspects of the full many-body problem. Density- 87 References [1] Proceedings of the VIIM'th Brazilian Summer School on Electronic Structure (2002), eds. H. P. Santos, P, Z. Coura, S, O. Dantas and P.M. V. B. Barone. 2] J. Tho, J. P. Perdow, V. N. Staroverov and G. B. Scuseria, Phys. Rev. Lett. 91, 146401 (2003); J. Chem. Phys. 119, 12129 (2003); J. Chem. Phys, 120, 6808 (2004); Phys. Rev. B 69, 075102 (2004). [3] W. Kolm, Rev. Mod. Phys. 71, 1253 (1999). [4] J. A. Pople, Rev. Mod. Phys. 71, 1267 (1999) (5) R. M. Dreizler and E. K. U. Gross, Density Functional Theory (Springer, Berlin, 1990). [6] R. G. Par and W. Yang, Density-Functional Theory of Atoms and Molecules (Oxford University Press, Oxford, 1989) [7] W. Koch and M. C. Holthausen, A Chemist's Guide to Density Functional Theory (John Wiley & Sons, New York, 2001). {8} R. O. Jones and O. Gunnarsson, Rev. Mod. Phys. 61, 689 (1989). [9] J. M. Seminario (Ed.), Recent Developments end Applications of Modern DFT (Blsovier, Amsterdam, 1996). [10] R. F, Nalewajski (Hd.), Density Punctional Theory I-IV (Springer, ‘Top- jes in Current Chemistry Vols, 180-183, 1996) [11] V. 1. Anisimov (Ed.), Strong Coulomb Correlations in Blectromic Struc- ture Calculations: Beyond the Local Density Appromimation (Gordon & Breach, 1999). [12] N. H. March, Blectron Density Theory of Atoms and Molecules (Aca- demic Press, London, 1992). [13] B. B. Laird, R. B. Ross and. Ziegler (Eds.), Chemical Applications of Density Functional Theory, (American Cliemical Society, 1996). [14] D. P. Chong (Ba.), Recent Advances in Density Punctional Methods (Singapore, World Scientific, 1995) 59) [15] D. Joulbert (Ed), Density Functionals: Theory and Applications (Gpringer Lecture Notes in Physics Vol. 500, 1998) [16] C. Fiolhais, F. Nogueira and M. Marques (Bds.), A Primer in Density Functional Theory (Springer Lecture Notes in Physics Vol. 620, 2003) [17] J. F, Dobson, G. Vignale and M. P, Das (Bds.), Density Functional ‘Theory: Recent Progress and New Directions (Plenum, New York, 1998) [18] H. Bschrig, The Fundamentals of Density Functional Theory (Teubner, Leipzig, 1996). 19] B. K. U. Gross and R. M. Dreizler (Eds), Density Functional Theory (Plenum, New York, 1998) (20] N. Argaman and G. Makov, Am. J. Phys. 68, 69 (2000). [21] B.N. Beonomou, Green’s Functions in Quantum Physics (Springer, New York, 1979). [22] M. Levy, Phys. Rev. A 26, 1200 (1982). [23] B. H. Lieb in Density Functional Methods in Physics, edited by R. M, Dreizler and J. da Providencia, (Plenum, New York, 1985). [24] P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964). [25] B. Schrédinger, Ann, Physik 79, 361, 489, 734 (1926); 80, 437 (1926); 81, 109 (1926), [26] A. K. Theophilou, J. Phys. C 12, 5419 (1979). [27] 0. Gritsenko and B. J. Baerends, J. Chem. Phys. 120, 8364 (2004). [28] P. Strange, Relativistic Quantum Mechanics with Applications in Con- densed Matter and Atomic Physics (Cambridge University Press, Cam- bridge, 1998). {29} ‘P. Kato, Commun. Pure Appl. Math, 10, 151 (1957). [30] N. H. March, Self-consistent fields in atoms (Pergamon Press, Oxford, 1975) {31] J.P. Perdow and M. Levy, Phys. Rev. B 31, 6264 (1985) [82] HL. Bschrig and W. B. Pickett, Solid State Commun. 118, 123 (2001). [33] KK. Capelle and G. Vignale, Phys. Rev, Lett. 86, 5546 (2001). {34] K. Capello and G. Vignale, Phys. Rev. B 65, 113106 (2002). [35] IN. Argaman and G. Makov, Phys. Rev. B 66, 052418 (2002). [86] N. Gidopolous in The fundamentals of density matriz and density func- tional theory in atoms, molecules and solids, N. Gidopoulos and $. Wilson cds. (Kluwer Series ‘Progress in Theoretical Chemistry and Physics’, 2003); and cond-mat /0510199. [37] W. Kohn, A. Savin and G. A, Ullrich, Int. J. Quantum Chem. 100, 20 (2004). [38] C. A. Ullrich, Phys. Rev. B 72, 073102 (2005) (39) ‘T. L. Gilbert, Phys. Rev. B 12, 2111 (1975). (40) J. E. Harriman, Phys. Rev. A 24, 680 (1981). [i] J.T. Chayes, L. Chayes and M. B. Ruskai, J. Stat. Phys. 38, 497 (1985) [42] C. A. Ulltich and W. Kohn, Phys. Rev. Lett, 89, 156401 (2002); ibid 87, 093001 (2001) [43] P. E. Lammert, J. Chem. Phys. 125, 074114 (2006). [44] A. P. Favaro, J. V. Batista Ferreira and K, Capelle, Phys. Rev. B 73, 045133 (2006) (45] L. W. Wang and M. P. Teter, Phys. Rev. B 45, 13196 (1992). [46] M. Foley and P. A. Madden, Phys. Rev. B 53, 10589 (1996). [47] B. J. Zhou, V. L. Ligneres and B. A. Carter, J. Chem, Phys. 122, 044103 (2005) 48] EB. K. U. Gross, E. Runge and O. Heinonen, Many Particle Theory (Adam Hilger, 1992) 61 [49] A. Szabo and N. 8. Ostlund, Modern Quantum Chemistry (McGraw- Hill, New York, 1989). (50) L. J. Sham and M, Sebliiter, Phys, Rev. Lett. 51, 1888 (1983) [61] EB. Runge and B. K. U. Gross, Phys. Rev. Lett. 52, 997 (1984). {62] B. K. U. Gross, M. Petersilka and T. Grabo in Ref, (13) [63] P. Nozieres and D. Pinos, The Theory of Quantum Liquids (Addison Wesley, New York, revised printing 1989) 54] N. D. Mermin, Phys. Rev. 187, 1441 (1965). [58] M. Levy and J. P. Perdew, Phys. Rev. A 32, 2010 (1985). 56] B, H. Lieb and 8. Oxford, Int, J. Quantum Chem. 19, 427 (1981). 57] G. K-L. Chan and N. C. Handy, Phys. Rev. A 59, 3075 (1999). {58} J. P. Perdew, R. G. Parr, M. Levy and J. L. Balduz, Phys. Rev. Lett. 49, 1691 (1982) [59] J. P, Perdew and M. Levy, Phys. Rev. Lett. 51, 1884 (1983). [60] W. Kohn and L. J. Sham, Phys, Rev. 140, A133 (1965) [61] W. Kohn, A. D. Becke and R. G. Parr, J. Phys. Chem. 100, 12974 (1996). (62} C. 0. Almbladh and U. v. Barth, Phys. Rev. B 81, 3231 (1985). [63] M. Levy, J. P. Pordew and V. Sahni, Phys. Rev. A 30, 2745 (1984) [64] J. Chen, J. B, Krieger, Y. Li and G. J, Iafrate, Phys. Rev. A 54, 3939 (1996). {65] M. Liiders, A. Ernst, W. M. ‘Temmerman, %. Szotek and P. J. Durham, J. Phys. Cond. Mat, 13, 8587 (2001). (66] J. Muskat, A. Wander, and N. M. Harrison, Chem, Phys. Lett. 842, 397 (2001). 62. (67] A. Savin, C. J. Umrigar and X. Gonze, Chem. Phys, Lett, 288, 391 (1998). " 1998). (68) W. G. Aulbur, L. Jénsson and J. W. Wilkins, Solid State Phys. 54, 1 (1999). [69] F. Aryasetiawan and O. Gunnarsson, Rep. Prog. Phys. 61, 237 (1998). [70] G. Onida, L. Reining and A, Rubio, Rev. Mod. Phys. 74, 601 (2002). [71] A. Svane and 0. Gunnarsson, Phys. Rev. Lett. 65, 1148 (1990); abid'72, ‘1248 (1994) {72] Z. Szotek, W. M. ‘Temmermann and H. Winter, Phys. Rev. Lett. 72, 1244 (1994). [73] P. Strange, A. Svane, W. M. ‘Temmermann, Z, Seotek and H. Winter, Nature 399, 756 (1999). (74] M. R. Norman and D. D. Koelling, Phys. Rev. B 30, 5530 (1984). [75] 0. K. Andersen, Phys. Rev. B 12, 3060 (1975) [76] 0. K. Andersen, O. Jepsen and D, Gldtzel in Highlights of Condensed ‘Matter Theory, P. Bassani, F. Funi and M. P. Tosi, eds. (North Holland, Amsterdam, 1985). [?'7] 0. K. Andersen and 'T, Saha-Dasgupta, Phys. Rev. B 62, 16219 (2000). {78] B. Zurck, O. Jepsen and O. K. Andersen, Chem.Phys.Chem. 6, 1934 (20085) {79} T. H, Dunning Jr., J. Chem. Phys. 90, 1007 (1989) [80] A. B. F. da Silva, Hl. F, M. da Costa and M. Trsie, Mol. Phys. 68, 433 (1989) [81] P. B. Joxge, B. V. R. de Castro and A. B. F. da Silva, J. Comp. Chem. 18, 1565 (1997). [82] G. B. Bachelet, D. R. Hamann and M. Schkiter, Phys. Rev. B 25, 2103 (1982), {83] L. Kicinman and D. M. Bylander, Phys. Rev. Lett. 48, 1425 (1982). [84] D. Vanderbilt, Phys. Rev. B 41, 7802 (1990). {85] N. Troullier and J. L. Martins, Phys. Rev. B 43, 1993 (1991). [86] W. B, Pickett, Comp. Phys. Rep. 9, 116 (1989) [87] D. Porezag, M. R. Pederson and A. Y. Liu, Phys. Stat, Sol. (b) 217, 219 (2000) [88] M. C. Payne, M. P. ‘Teter, D. C. Allan, T. A. Arias and J. D. Joannopou- Jos, Rev. Mod. Phys. 64, 1045 (1992). [89] U. von Barth and 1. Hedin, J. Phys. C5, 1629 (1972). [90] 0. Gunnarsson and B. Lundavist, Phys. Rev. B 13, 4274 (1976). (91) D. M. Geperley and B. J. Alder, Phys. Rev. Lett. 45, 566 (1980). [92] S. H. Vosko, L. Wilk and M. Nusair, Can. J. Phys. 58, 1200 (1980). {93] J. P, Perdew and A. Zunger, Phys. Rev. B 23, 5048 (1981). [94] J. P. Perdew and ¥. Wang, Phys. Rev. B 45, 18244 (1992). [95] J. P. Perdew and Y. Wang, Phys. Rev. B 33, 8800 (1986) [96] J. P. Perdew, K. Burke and M. Emzerhof, Phys. Rev. Lett. 77, 3865 (1996); ibid 78, 1396(B) (1997). See also V. Staroverov et al., Phys. Rev. A 74, 044501 (2006). [97] A. D. Becke, Phys. Rev. A 38, 3098 (1988). [98] C. Leo, W. Yang and R. G. Parr, Phys. Rev. B 37, 785 (1988). [99] D. C. Patton and M. R. Pederson, Phys. Rev. A 56, 2495 (1997). [100] J. Tao and J. P. Perdew, J. Chem. Phys. 122, 114102 (2005). {101} W. Kohn, Y. Meir and D. E, Makarov, Phys. Rev. Lett. 80, 4153 (1998). of [102] M. Lein, J. F. Dobson and BE. K. U. Gross, J. Comp, Chem. 20, 12 (1999), [103] J. F. Dobson and B. P. Dinte, Phys, Rev. Lett, 76, 1780 (1996). [104] Y. Andersson, D. C. Langreth and B. I, Lundqvist, Phys. Rev. Lett. 76, 102 (1996). [195] B. R. Johnson and A. D. Becke, J. Chem. Phys. 128, 024101 (2005). [206] G. Filippi, ©. J. Umrigar and M, ‘Taut, J. Chem, Phys. 100, 1290 (1994). [107] P. Ziesche, S. Kurth and J. P. Perdew, Comp. Mat. Sci. 11, 122 (1998) [108] A. D. Becke, J. Chem. Phys. 98, 5648 (1993). [109] A. D. Becke, J. Chem. Phys. 107, 8554 (1997). See also A. D. Becke, J. Comp. Chem. 20, 63 (1999). [110] D. J, Tozer and N. C. Handy, J. Chom. Phys. 108, 2545 (1998) [111] ‘7. van Voorhis and G. B, Scuseria, J. Chem. Phys. 109, 400 (1998) [112] J. P. Perdew, S. Kurth, A. Zupan and P. Blaha, Phys. Rey. Lett. 82, 2544 (1999) [113] A. D. Becke, J. Chem. Phys. 104, 1040 (1996). [114) Y. H. Ki, 1. H. Lee, 8. Nagaraja, J. P. Leburton, R. Q. Hood, and R.M, Martin, Phys. Rev. B 61, 5202 (2000). [115] S. Kurth, J. P. Perdew and P. Blaha, Int. J. Quantum Chem. 75, 889 (1999). {116] C. Adamo, M. Brnzerhof and G. B. Scuseria, J. Chem Phys. 112, 2643 (2000). [117] 0. A. Vydrow and G. B, Seuseria, J. Chem, Phys. 121, 8187 (2004) [118] U. Lundin and O. Eriksson, Int, J. Quantum, Chem. 81, 247 (2001), {119} C. Legrand, B, Suraud and P.-G. Reinhard, J. Phys. B $5, 1115 (2002) 65 [120] J. B. Krieger, ¥. Li and G. J. Iafrate, Phys, Rev. A 45, 101 (1992); ibid 46, 5453 (1992); abid 47 165 (1993). (121) B. Engel and S. H. Vosko, Phys. Rev. A 47, 2800 (1993). [122] 'T. Grabo, T. Kreibich, 8. Kurth and E. K, U, Gross, in Ref. [11] [123] T. Grabo and E, K. U. Gross, Int. J. Quantum Chem. 64, 95 (1997) T. Grabo and B. K. U. Gross, Chem, Phys, Lett, 240, 141 (1995). [124] T. Kotani, Phys. Rev. Lett. 74, 2989 (1995), [125] M. Stadele, J. A. Majewski, P. Vogl and A. Gérling, Phys. Rev. Lett. 79, 2080 (1997). M. Stadele, M. Moukara, J. A. Majewski, P. Vogl and A. Gérling, Phys. Rev. B 59, 10031 (1999). Y.-H. Kim, M. Stadele and R. M. Martin, Phys. Rev. A 60, 3633 (1999). [126] $. Kimmel and J. P. Perdew, Phys. Rev. B 68, 035103 (2003). [127] W. Yang and Q. Qu, Phys. Rev. Lett, 89, 148002 (2002). [128] V. N. Staroverov, G. B, Scuseria and B. R. Davidson, J. Chem. Phys. 125, 081104 (2006). [129] 0. Gunnarsson, M. Jonson and B. I. Lundqvist, Phys. Rev. B 20, 3136 (1979). [130] A. Cancio, M. Y. Chou and R. 0. Hood, Phys. Rev. B 64, 115112 (2002). [131] P. Garcia-Gonzaler, Phys. Rev. B 62, 2321 (2000). [132] G. L. Oliver and J. P. Perdew, Phys. Rev. A 20, 397 (1979). [183] K. Capelle and V. 1. Libero, Int. J. Quantum Chem. 105, 679 (2008). [134] L. M. Sandratsldi, Adv. Phys. 47, 91 (1998). (135] L. Nordstrdm and D. Singh, Phys. Rov. Lett. 76, 4420 (1996). {136] K. Capelle and 1. N. Oliveira, Phys. Rev. B 61, 15228 (2000). [137] A. K. Rajagopal and J. Callaway, Phys. Rev. B 7, 1912 (1973) 66 | [138] A. H. Mac Donald and 8. H. Vosko, J. Phys. C 12, 2977 (1979). [139] E. Bagel, T. Auth and R. M, Dreizler, Phys. Rev. B 64, 235126 (2001). [140] G. Vignale and M, Rasolt, Phys. Rev. Lett. 59, 2360 (1987). [141] G. Vignale and M. Rasolt, Phys. Rev. B 37, 10685 (1988). [142] M. Ferconi and G. Vignale, Phys. Rev. B 50, 14722 (1997). [143] M. Battocletti, H. bert and B, K. U, Gross, Europhys, Lett. 40, 45 (1997). [144] A. M. Lee, N. C. Handy and $, M, Colwell, J. Chem. Phys. 103, 10095 (1996). [145] K. Capelle, Phys. Rev. B 65, 100515 (2002). [146] K. Capelle and E. K. U. Gross, Phys. Rev. Lett. 78, 1872 (1997). [147] P. Skudlarski and G. Vignale, Phys. Rev. B 48, 8547 (1993), [148] B. Orestes, T. Marcasso and K, Capelle, Phys. Rev. A 68, 022105 (2008). [149] L. N. Oliveira, HB. K. U. Gross and W. Kohn, Phys. Rev. Lett. 60, 2430 (1988). [150] W. Koha, B. K. U. Gross and I. N. Oliveira, J. de Physique 50, 2601 (1989). [151] W. M, Temmermann, 2. Szotek, B. L. Gyorffy, 0. K. Andersen, and . Jepsen, Phys. Rev. Lett. 76, 307 (1996) [152] 8. Kurth, M. Marques, M. Liiders and E. K. U. Gross, Phys, Rev. Lett. 83, 2628 (1999) [153] K. Capelle and LN. Oliveira, Europhys. Lett. 49, 376 (2000). [154] U. von Barth, Phys. Rev. A 20, 1693 (1979) [155] A. Gérling, Phys. Rev. Lett. 85, 4229 (2000). 67 | [156] B. K. U. Gross, L. N. Oliveira and W. Kohn, Phys. Rev. A 37, 2805; ibid 2809; ibid 2821 (1988), [157] A. Nagy, Phys. Rev. A 49, 3074 (1994); ibid 42, 4388 (1990). [158] A. Gérling, Phys. Rev A 59, 3359 (1999). A. Gérling, Phys. Rev A 54, 3912 (1996) {159] A. Nagy, Chem. Phys. Lett. 296, 489 (1998). M. Levy and A. Nagy, Phys. Rev. Lett. 83, 4361 (1999). (160] K. Capelle, J. Chem. Phys. 119, 1285 (2003) [161] R. van Leeuwen, Phys. Rev. Lett. 82, 3863 (1999). (162) B. K. U. Gross, J. F. Dobson and M. Petersilka in Ref. (10). K, Burke and B. K, U, Gross in Ref. [15]. [163] M. Petersilka, U. J. Gossmann and B. K. U. Gross, Phys. Rev. Lett. 76, 1212 (1996). See also 'T. Grabo, M. Petersilka and B. K. U. Gross, J Mol. Struc. (Theochem) 501, 353 (2000) [164] M. E. Casida, in Ref. (14). J. Jamorski, M. BE. Casida and D. R. Salahub, J. Chem, Phys. 104, 5134 (1996), [165] S. J. A. van Gisbergen, J. G. Snijders, G. te Velde and B. J. Baerends, Comp. Phys. Comm. 118, 119 (1999). [166] R. 8, Stratmann, G. B. Scuseria and M. J. Frisch, J. Chem, Phys. 109, 8218 (1998) [167] M. Lein, B. K. U, Gross and V. Engel, Phys. Rev. Lett. 85, 4707 (2000). [168] C. A. Ullrich and B. K. U. Gross, Comm. At, Mol, Phys. 33, 211 (1997). {169} K. Capelle, G. Vignale and B. L. Gyorfly, Phys. Rev. Lett. 87, 206403, (2001). K. Capelle and B. L. Gyorffy, Burophys. Lett. 61, 354 (2003). [170] 0, Gunnarsson and K. Schénhammer, Phys. Rev. Lett. 56, 1968 (1986). 68 [171] N. A. Lima, M. F. Silva, L. N. Oliveira and K. Capelle, Phys. Rev. Lott. 90, 146402 (2003). N. A. Lima, L. N. Oliveira and K. Capelle, Bu- rophys. Lett. 60, 601 (2002). M. F. Silva, N. A. Lima, A. L, Malvezzi and K, Capelle, Phys. Rev. B 71, 125130 (2005). G, Xianlong et al., Phys. Rev. B 78, 165120 (2008) [172] R. J. Magyar and K. Burke, Phys. Rev. A 70, 032508 (2004), [173] V. L. Libero and K. Capelle, Phys. Rev. B 68, 024423 (2003) and Phys, B 384, 179 (2006). P. E. G. Assis, V. L. Libero and K. Capelle, Phys. Rev. B 71, 052402 (2005). See also Ref. [133] [174] V. L. Libero and K. Capelle, cond-mat/0506206. [175] H. Chermette, J. Comp. Chem. 20, 129 (1999). [176] F. De Proft and P. Geerlings, Chem. Rev. 101, 1451 (2001). [177] A. B. Mattsson, Science 298, 759 (2002), 6

You might also like